diff --git "a/spin dynamics/1.json" "b/spin dynamics/1.json" new file mode 100644--- /dev/null +++ "b/spin dynamics/1.json" @@ -0,0 +1 @@ +[ { "title": "1609.05603v1.Effective_indirect_multi_site_spin_spin_interactions_in_the_s_d_f__model.pdf", "content": "arXiv:1609.05603v1 [cond-mat.str-el] 19 Sep 2016Effective indirect multi-site spin-spin interactions\nin the s-d(f) model\nK.K.Komarova), D.M.Dzebisashvilia,b)\naKirensky Institute of Physics, Federal Research Center KSC SB RAS, 660036, Krasnoyarsk,\nRussian Federation\nbSiberian State Aerospace University, Krasnoyarsk, 660014 , Russian Federation\nUsing the diagram technique for Matsubara Green’s function s it is shown that the dynamics\nof the localized spin subsystem in the s-d(f) model can be des cribed in terms of an effective spin\nmodel with multi-site spin-spin interactions. An exact rep resentation of the action for the effective\npurely spin model is derived as an infinite series in powers of s-d(f) exchange interaction J. The\nindirect interactions of the 2-nd, 3-rd and 4-th order are di scussed.\n1. Introduction\nKondo lattice model or s-d(f) exchange model is widely used to desc ribe the correlation\neffects in metals (and their compounds) with unfilled d- andf-shells.\nThe nature of the ground state in these systems is largely determin ed by the result of\ncompetition between two interactions. On the one hand, s-d(f) ex change coupling between\nspins of itinerant s- and localized d(f)- electrons due to Kondo fluct uations screens the\nlocalized spins and tends to form a non-magnetic ground state [1]. I n the opposite direction\nacts indirect exchange (RKKY) interaction between spins of d(f)- electrons [2], trying to set\na long-rang magnetic order which is not necessarily a ferromagnetic or antiferromagnetic.\nCorrelation effects can lead to, for example, helical magnetic struc tures [3], with period that\nis determined by the Fermi surface singularities [5, 6].\nIt is clear that the study of competition between different effective interactions, that\noccur in the localized spin subsystem and are caused by the same itine rant electrons, must\nbe carried out within the same unified approach.\nThe purpose of this paper is to derive such an effective Hamiltonian (o r rather action),\nwhich will allow to study the localized spin subsystem in the s-d(f) mode l within the model\nwith only spin-spin interactions. At the same time all possible spin-spin interactions that\noccur in different orders of perturbation theory in J, should be determined by the same\nmodel parameters: 1) the s-d(f)-exchange coupling constant J; 2) the dispersion law of\nconduction electrons; 3) the filling degree of the conduction band. Such an approach should\nallow to analyze all types of exchange spin-spin interactions from a u nified point of view.\nThis problem is solved by integrating charge degrees of freedom usin g the diagram tech-\nnique for Matsubara Green’s functions. It is shown that in addition t o the two-spin indirect\nexchange interaction, which occurs in the second order in s-d(f)- parameter J, in the follow-\ning orders also appear terms describing, in particular, the ring and b iquadratic exchange\ninteractions. The essential point of all these interactions is the ac count for retardation ef-\nfects provided by the imaginary time dependence of all the effective multi-site exchange\nparameters.\n12. The Hamiltonian of the s-d(f) exchange model\nThe Hamiltonian of the Kondo lattice model (or s-d(f) exchange mod el) can be written as a\nsum of two terms:\nˆH=ˆH0+ˆHint, (1)\nwhere\nˆH0=/summationdisplay\nkα(εk−µ)c+\nkαckα,ˆHint=J\n2/summationdisplay\nfc+\nf˜Sfcf. (2)\nOperator ˆH0stands for the energy of noninteracting current carriers (elect rons or holes) with\ndispersion εk,µ– chemical potential. Operator c+\nkα(ckα) creates (annihilates) a particle in\nthe state with quasimomentum kand spin projection α=±1/2.\nThe second term in (1) describes Kondo exchange interaction betw een localized spins\nand itinerant quasiparticles. The intensity of this interaction is defin ed by the constant\nJ. Operator ˜Sf=/vectorSf/vector σis a product of a localized spin vector operator /vectorSf= (Sx\nf,Sy\nf,Sz\nf)\nand a vector /vector σ= (σx,σy,σz) which is formed of Pauli matrices. In definition (2) the spinor\nnotations:\nc+\nf=/parenleftbigc+\nf↑c+\nf↓/parenrightbig\n, cf=/parenleftbiggcf↑\ncf↓/parenrightbigg\n,˜Sf=/parenleftbiggSz\nfS−\nf\nS+\nf−Sz\nf/parenrightbigg\n, S±\nf=Sx\nf±iSy\nf, (3)\nare used. The operators cfandckare related to each other by Fourier transformation:\ncf=N−1/2/summationtext\nkeikfck.\n3. Spin Green’s functions and effective action\nThe thermodynamic properties of localized spin subsystem convenie ntly studied on the basis\nof the diagram technique for Matsubara Green’s functions [7]:\nGjj′(f−f′,τ−τ′) =−/angbracketleftBig\nTτ¯Sj\nf(τ)¯Sj′\nf′(τ′)/angbracketrightBig\n, j,j′={x,y,z}. (4)\nIn this expression the spin operators are written down in the Heisen berg representation:\n¯Sj\nf(τ) =eτˆHSj\nfe−τˆH, where ˆHis the s-d(f) exchange Hamiltonian (1), and τ– imaginary\ntime varying within interval (0 ,1/T) (T– is the temperature). The imaginary time ordering\noperator Tτarranges all operators on the right side of it in the descending orde r of the\nimaginary time τfrom left to right. Angle brackets in (4) denote a thermodynamic av erage\nover the grand canonical ensemble described by the Hamiltonian ˆH.\nAs is known [8], in the interaction representation: Sj\nf(τ) =eτˆH0Sj\nfe−τˆH0, the expression\n(4) transforms into:\nGjj′(f−f′,τ−τ′) =−/angbracketleftBig\nTτSj\nf(τ)Sj′\nf′(τ′)S(β)/angbracketrightBig\n0. (5)\nHere the scattering matrix S(β) is defined by the formula:\nS(β) =Tτexp/braceleftbigg\n−/integraldisplayβ\n0dτˆHint(τ)/bracerightbigg\n, (6)\n2whereintheexponentundertheintegralstandstheinteractiono perator(2)intheinteraction\nrepresentation. Lower index ”0”, on the right of the angle bracke ts in (5) indicates that the\nthermodynamic averaging is over the ensemble of non-interacting s pin and fermion systems.\nBesides, when expanding the thermodynamic averages using Wick’s t heorem only connected\ndiagrams should be considered.\nNote that for calculating the Green’s function (5) along with the usu al diagramtechnique\n[8] it is necessary to use the diagram technique for spin operators [9 ] as well. The expanding\nof aTτ-ordered average of a product of spin S- and fermion c-operators can be divided into\ntwo stages: first, only c-operators are paired, and then the Wick’s theorem is applied to the\nremaining spin operators. Formally this can be written down as follows :\n/angbracketleftbig\nTτS1...Slc1...c+\nm/angbracketrightbig\n0=/angbracketleftBig\nTτS1...Sl/angbracketleftbig\nTτc1...c+\nm/angbracketrightbig\nc0/angbracketrightBig\nS0. (7)\nInternal thermodynamic averageontherightsideoftheequation (7)withindex ” c0”denotes\naveraging only over an ensemble of non-interacting fermions. Exte rnalTτ-ordered averaging,\nindicated by the index ” S0”, should be done using the ensemble of non-interacting spin\nsubsystems. Applying equation (7) to the definition of the Green’s f unction (5) we can\nwrite:\nGjj′(f−f′,τ−τ′) =−/angbracketleftBig\nTτSj\nf(τ)Sj′\nf′(τ′)SS(β)/angbracketrightBig\nS0, (8)\nwhere the effective scattering matrix SS(β) is defined by the expression:\nSS(β) =/an}bracketle{tS(β)/an}bracketri}htc0. (9)\nTo obtaintheeffective purely spin model first weshould pair all the c-operatorsaccording\nto the Wick’s theorem in each order of the S-matrix expansion in powers of the coupling\nconstants J. Afterthattheresultmustberewrittenasa Tτ-orderedexponent. Theargument\nof the exponent will determine the required effective interaction in t he subsystem of localized\nspins. Note that the described here scheme of integrating over th e charge degrees of freedom\nin the s-d(f) model in some sense is similar to the proof of equivalence between diagrammatic\nexpansion for the Green’s function of localized f-electrons in the periodic Anderson model\nand diagrammatic expansion of the fermion Green’s function in the Hu bbard model [10, 11].\nLet us consider for the average /an}bracketle{tS(β)/an}bracketri}htc0the diagrams of n-th order with respect to\nthe constant J. Since each term in ˆHintconsists of one spin operator and two second-\nquantization operators ( c+andc), the structure of diagrams emerging after all possible\npairings of c-operators is characterized by a some number of fermionic loops (s ee figure\n1). The order of each loop is determined by the number of fermionic lin es, corresponding\nto the fermion Green’s functions G(0)\nk(τ−τ′) =−/an}bracketle{tTτck(τ)c+\nk(τ′)/an}bracketri}ht0, and the same number of\nvertices. In contrast to the conventional diagram technique in th is case each vertex is related\nnot to a function (or a number) but to the spin operator. Therefo re then-th order diagram\ncorresponds to a product of nspin operators entering to the Tτ-ordered thermodynamic\naverage/an}bracketle{tTτ.../an}bracketri}htS0.\nLet us denote by pjthe number of loops of the same order mj, where the index j\nenumerates all loops of different order. Then obviously n=p1m1+p2m2+.... Until the\nspin operators are not paired the analytical expression correspo nding to each n-th order\ndiagram can be represented as a product of some multipliers. Each m ultiplier is related to\na single fermionic loop. Note that in the case of further application of the spin diagram\n3L\n1pjp121m3\n121m3\n12jm3\n12jm3\nLL L\nFigure 1: The n-th order diagrams, arising due to pairing of all c-operators in the expansion\nof/an}bracketle{tS(β)/an}bracketri}htc0, can be represented as a set of fermionic loops. The order of each loopmjis\ndetermined by the number of vertices represented by points and t he same number of lines\nwith arrows denoting the fermion propagator. Each vertex corre sponds to the spin operator.\nIf the number of vertices of the same order mjis denoted by pj, then:p1m1+p2m2+···=n.\ntechnique [9] there should be taken into account only such pairings o f spin operators when\nall fermionic loops are connected graphically with elements of the spin diagram technique.\nThe fermion loop of the mj-th order is formed due to mjoperators ˆHint. Let’s use the\nindex ”L” in angle brackets /an}bracketle{t.../an}bracketri}htLto indicate the fact that when pairing all the c-operators\nin theTτ-ordered average of mjoperators ˆHintthere should be taken into account only one-\nloop diagrams of mj-th order. Then the analytical contribution to the one loop of mj-th\norder will be determined by the average: /an}bracketle{tTτˆHint(τ1)...ˆHint(τmj)/an}bracketri}htLc0. Strictly speaking this\naverage gives rise to ( mj−1)! equivalent fermionic loops of mj-th order. Using the definition\nof the average /an}bracketle{tTτ.../an}bracketri}htLc0the expansion of /an}bracketle{tS(β)/an}bracketri}htc0can be written as:\n/angbracketleftbig\nS(β)/angbracketrightbig\nc0=Tτ/summationdisplay\nn(−1)n\nn!/summationdisplay\np1,p2...\nm1,m2...\n(p1m1+···=n)A(p1m1;p2m2;...)×\n×/productdisplay\nj/integraldisplay\ndτ(j)\n1...dτ(j)\nm1/angbracketleftbig\nTτˆHint(τ(j)\n1)...ˆHint(τ(j)\nm1)/angbracketrightbig\nLc0. (10)\nInderivationtheformula(10)ineach n-thorderof ˆHintthethermodynamicaverage /an}bracketle{tTτ.../an}bracketri}htc0\nwas represented as a sum of products of averages /an}bracketle{tTτ.../an}bracketri}htLc0. The second sum in (10) takes\ninto account all possible partitions of noperators ˆHintintop1+p2+...groups. Each\ngroup corresponds to a single fermionic loop and the equality: p1m1+p2m2+···=nmust\nbe satisfied. The index jin the product/producttext\njruns all groups (loops) at a fixed partition.\nBecause the averaging in the /an}bracketle{tS(β)/an}bracketri}htc0is carried out only over the charge degrees of freedom,\nthe remaining spin operators must be still ordered by the imaginary t ime. This justifies the\npresence of operator Tτin the right side of (10). Combinatorial factor A(p1m1;p2m2;...)\ntakes into account the equivalent contributions arising at each pos sible partitioning of n\noperators ˆHintin groups.\nIt can be shown (cf. [8]) that:\nA(p1m1;p2m2;...) =n!\np1!(m1!)p1p2!(m2!)p2.... (11)\nIntroducing notations:\nΞm=(−1)m+1\nm!/integraldisplayβ\n0dτ1...dτm/angbracketleftBig\nTτˆHint(τ1)...ˆHint(τm)/angbracketrightBig\nLc0, (12)\n4Figure 2: The effective action Ξ can be expressed as an infinite series of terms Ξ n. Each\nterm Ξ nin the diagrammatic representation corresponds to a single loop of n-th order in the\nexchange coupling constant Jand describes the effective n-site spin-spin interaction. The\nlines with arrows represent bare fermion Green’s functions and eac h vertex, shown with a\nbold circle, corresponds to a spin operator ˜S.\nthe expression (10) can be rewritten as:\n/angbracketleftbig\nS(β)/angbracketrightbig\nc0≡SS(β) =Tτ/summationdisplay\np1,p2...1\np1!(−Ξ1)p11\np2!(−Ξ2)p2···=Tτexp{−Ξ1−Ξ2−...}.(13)\nUsing the explicit form of ˆHintcarry out in (12) pairings of all the c-operators. Then\nintroducing the variable xj= (/vectorRfj,τj) we obtain expression for the operator Ξ min the\nWannier representation:\nΞm=1\nm/parenleftbiggJ\n2/parenrightbiggm/integraldisplay\ndx1...dxmG(0)(x1−x2)...G(0)(xm−x1) Sp/braceleftBig\n˜S(x1)...˜S(xm)/bracerightBig\n,(14)\nwhere the integral over d xdenotes the operation/integraltextβ\n0dτ/summationtext\nfand the trace is calculated over\nthe Pauli matrix indices.\nFrom the formula (13) it follows that the effective interactions in the localized spin\nsubsystem are determined by the structure of the series:\nΞ =∞/summationdisplay\nm=1Ξm. (15)\nThe effective action Ξ describes all possible multi-site spin-spin intera ctions in the localized\nspin subsystem in the arbitrarily order of the coupling constant J. The partial action Ξ m\ndetermines the m-th order interactions and diagrammatically can be represented as a loop\nwithmlines, corresponding to propagators G(0), andmvertices related to spin operators ˜S\n(see figure 2). It can be seen that all effective interactions in Ξ tak e into account retardation\neffects. This is due to imaginary time dependence of all the effective in teractions in each\nterm (14) of the series for Ξ.\n4. Effective interactions of the 2nd, 3rd and 4th order\ninJ\nLet us consider the first several terms of the series for Ξ. The fir st term with n= 1 is zero,\nbecause Sp {˜S(x1)}=/summationtext\njSj(x1)Sp{σj}, and Sp{σj}= 0 at any j=x,y,z.\n5To calculate the operator Ξ 2one should to derive the trace Sp/braceleftBig\n˜S(x1)˜S(x2)/bracerightBig\n. Using the\nidentity: σiσj=δij+iεijlσl, whereεijl– Levi-Civita antisymmetric tensor we find:\nSp/braceleftBig\n˜S(x1)˜S(x2)/bracerightBig\n= 2/parenleftBig\n/vectorS(x1)/vectorS(x2)/parenrightBig\n. (16)\nThen the operator Ξ 2takes the form:\nΞ2=/integraldisplay\ndx1dx2V2(x1−x2)/vectorS(x1)/vectorS(x2), (17)\nwhere the effective interaction between spins is defined via a polariza tion loop (see figure 3):\nV2(x1−x2) =/parenleftbiggJ\n2/parenrightbigg2\nG(0)(x1−x2)G(0)(x2−x1). (18)\nFigure 3: The action Ξ 2can be represented by a loop of the second order in the exchange\ncoupling constant J.\nFrom equation (17) it follows, that at f1/ne}ationslash=f2the second order effective action Ξ 2de-\nscribesindirect exchangeinteractionoftwo localizedspinsthrough thesubsystem ofitinerant\nelectrons. However, in contrast to the usual RKKY-interaction h ere the retardation effects,\ncaused by τ-dependence of V2, are taken into account. Note that in the Ξ 2there is also a\nterm with f1=f2. The Fourier transform of the indirect exchange interaction (18) has the\nform:\nV2(k,iωm) =/parenleftbiggJ\n2/parenrightbigg2\nχ0(k,iωm), (19)\nwhere\nχ0(k,iωm) =1\nN/summationdisplay\nqfq−fq+k\niωm+εq−εq+k(20)\nis the Lindhard susceptibility in the Matsubara representation, ωm= 2mπTwithm∈Z,\nandfq= (exp{(εq−µ)/T}+1)−1is the Fermi-Dirac distribution function. The intensity of\nthe exchange interaction is largely determined by the properties of the itinerant subsystem.\nThe well known RKKY-interaction follows from (20) at ωm≡0.\nCalculating the 3rd order effective action Ξ 3we obtain (see also figure 4):\nΞ3=/integraldisplay\ndx1dx2dx3V3(x1,x2,x3)/vectorS(x1)·/parenleftBig\n/vectorS(x2)×/vectorS(x3)/parenrightBig\n, (21)\nwhere\n6Figure 4: Diagrammatic representation of the third order effective action Ξ 3.\nFigure 5: Fourth order effective action Ξ 4in the diagrammatic representation.\nV3(x1,x2,x3) =i2\n3/parenleftbiggJ\n2/parenrightbigg3\nG(0)(x1−x2)G(0)(x2−x3)G(0)(x3−x1). (22)\nThe formula (21) describes the three-spin interactions in the form of a mixed product of\nthree spin operators. This type of interaction was first considere d in [12] but without taking\ninto account the retardation effects.\nItisevidentthattheinteraction(21)favorsthechiralorderinth emagneticsubsystem. In\nthis regard we note that in the paper [13] the third order correctio ns to the Hall conductivity\ndue to s-d(f) exchange interaction were shown to give rise the ano malous Hall effect provided\nthat non trivial spin configuration (chirality) is formed in the spin sub system. Interestingly\nthe structure of the expression for the Hall conductivity obtaine d in [13] is similar to that\nof (21).\nThe operator Ξ 4, which determines the effective spin-spin interactions in the fourth order\nin the coupling constant J, after calculating the trace Sp {˜S(x1)˜S(x2)˜S(x3)˜S(x4)}takes the\nform (see also figure 5):\nΞ4=/integraldisplay\ndx1dx2dx3dx4V4(x1,x2,x3,x4)/parenleftBig\n/vectorS(x1)/vectorS(x2)/parenrightBig\n·/parenleftBig\n/vectorS(x3)/vectorS(x4)/parenrightBig\n, (23)\nwhere\nV4(x1,x2,x3,x4) =1\n2/parenleftbiggJ\n2/parenrightbigg4\n[G(0)(x1−x2)G(0)(x2−x3)G(0)(x3−x4)G(0)(x4−x1)−\n−G(0)(x1−x3)G(0)(x3−x2)G(0)(x2−x4)G(0)(x4−x1)+\n+G(0)(x1−x4)G(0)(x4−x3)G(0)(x3−x2)G(0)(x2−x1) ].(24)\nThe terms ofthe expression (23) with unequal indices of sites fj(j= 1,...,4) correspond\nto the four-spin exchange interactions. Among them there are, in particular, interactions\ndescribing ring exchange of four spins located, for example, at the square plaquette vertices.\nFor the first time the ring exchange (without retardation effects) was obtained from the\n7Hubbard model at half filling in the fourth order of the perturbation theory in the parameter\nt/U, wheretisthetunneling integral, and UistheCoulombrepulsion energyoftwo electrons\nat the same site [14]. The four-spin ring exchange interaction was inv olved to explain the\nmagnetic ordering features in the quantum crystal3He [15]. In the paper [16] it was argued\nthat ring exchange is important to describe magnetic properties of cuprate high-temperature\nsuperconductors. Effect of ring exchange interaction on the sup erconductivity in cuprates\nwas investigated in [17].\nAt pairwise coincident site indices: f3=f1andf4=f2, in the sum (23) there are terms\nthat are responsible for biquadratic exchange interaction. These interactions were first used\nin[18]forexplaining theparamagneticresonanceonMnionsinthecom poundMnO.Besides,\nthe biquadratic exchange interaction is essential in multilayer magne tic systems [19].\n5. Conclusion\nThe paper presents a method of deriving all possible kinds of effectiv e indirect interactions\nbetween the localizedspins dueto s-d(f) exchange coupling of thes e spins withthe subsystem\nof itinerant electrons. After integrating over charge degrees of freedom in the s-d(f) exchange\nmodel an exact representation for an action of a purely spin model is obtained. Using this\naction allows to study the spin subsystem in the s-d(f) model in the f ramework of effective\npurely spin model. The important point of this model is that all effectiv e interactions take\ninto account retardation effects. Although explicit expressions ar e written only for two-,\nthree- and four-spin interactions, the formula (14) allows to gene rate multi-site spin-spin\ninteractions in the arbitrary order in the s-d(f) exchange coupling constant J.\nThe work was supported by the Russian Foundation for Basic Resea rch (project nos.\n16-42-240435 and 15-42-04372).\nReferences\n[1] V.V. Moschalkov, N.B. Brandt, UFN 149(1986) 585.\n[2] M.A. Ruderman, C. Kittel, Phys. Rev. 96(1954) 99; K. Yosida, Phys. Rev. 106(1957)\n893; T. Kasuya, Prog. Theor. Phys. 16(1956) 45.\n[3] M. Hamada, H. Shimahara, Phys. Rev. B. 51(1995) 3027.\n[4] N. de C. Costa, J.P. de Lima, R.R. dos Santos, arXiv: 1506.00890v 1 [cond-mat.str-el]\n(2015).\n[5] K. Yosida, A. Watabe, Prog. Theor. Phys. 28(1961) 361.\n[6] I.E. Dzyaloshinskii, ZhETF 47(1964) 336.\n[7] T. Matsubara, Prog. Theor. Phys. 14(1955) 351.\n[8] A.A. Abrikosov, L.P. Gorkov, I.E. Dzyaloshinskiy, Methods of qua ntum field theory\nin statistical physics. – Moscow: Fizmatgiz, 1962. – 444 p.\n[9] Yu.A. Izyumov, F.A. Kassan-ogly, Yu.N. Skryabin, Field methods in the theory of\nferromagnetism. – Moscow: Nauka, 1974. S. – 224\n8[10] V.A. Moskalenko, TMF 110(1997) 308.\n[11] V.V. Val’kov, D.M. Dzebisashvili, Pis’ma v ZhETF 84(2006) 251.\n[12] H.A. Brown, JMMM 43(1984) L1-L2.\n[13] G. Tatara, H. Kawamura, JPSJ 71(2002) 2613.\n[14] M. Takahashi, J. Phys. C: Solid State Phys. 10(1977) 1289.\n[15] M. Roger, J.H. Hetherington, J.M. Delrieu, Rev. Mod. Phys. 55(1983) 1-64.\n[16] M. Roger, J.M. Delrieu, Phys. Rev. B 39(1989) 2299.\n[17] E.I. Shneyder, S.G. Ovchinnikov, A.V. Shnurenko, Pis’ma v ZhETF 95(2012) 211.\n[18] E.A.Harris, J.Owen, Phys. Rev. Lett. 11(1963) 9.\n[19] J.C. Slonczewski, J. Appl. Phys. 73(1993) 5957.\n9" }, { "title": "0906.5111v1.Steady_state_spin_densities_and_currents.pdf", "content": "arXiv:0906.5111v1 [cond-mat.mes-hall] 27 Jun 2009Steady-state spin densities and currents\nDimitrie Culcer1,2,3\n1Condensed Matter Theory Center, Department of Physics,\nUniversity of Maryland, College Park MD 20742.\n2Advanced Photon Source, Argonne National Laboratory, Argo nne, IL 60439.\n3Northern Illinois University, De Kalb, IL 60115.\n(Dated: December 4, 2018)\nThis article reviews steady-state spin densities and spin c urrents in materials with strong spin-\norbit interactions. These phenomena are intimately relate d to spin precession due to spin-orbit\ncoupling which has no equivalent in the steady state of charg e distributions. The focus will be\ninitially on effects originating from the band structure. In this case spin densities arise in an\nelectric field because a component of each spin is conserved d uring precession. Spin currents arise\nbecause a component of each spin is continually precessing. These two phenomena are due to\nindependent contributions to the steady-state density mat rix, and scattering between the conserved\nand precessing spin distributions has important consequen ces for spin dynamics and spin-related\neffects in general. In the latter part of the article extrinsi c effects such as skew scattering and side\njump will be discussed, and it will be shown that these effects are also modified considerably by\nspin precession. Theoretical and experimental progress in all areas will be reviewed.\nI. INTRODUCTION\nSpin electronics seeks to harness the spin degree of freedom of th e electron, aiming to generate and maintain a\nspin polarization for possible use in information storage through the manufacture of magnetic memory devices. For\nspin-based quantum computing the manipulation of a spin polarization is also crucial. A spin polarization can be\ngenegrated by optical, magnetic and electrical means. Optically, on e uses light to transfer angular momentum to\ncharge carriers during interband transitions.[1] This reliable proced ure has been tested successfully decades ago, yet\ndevices that require an input of light are currently impractical. Magn etically, one uses the Zeeman effect, to which\nsimilarobservationsapply. Electrically, onedrivesacurrentfromas pin-polarizedmaterialintoanunpolarizedoneand\nthus carry spin-polarized electrons across the interface. This wo rks in the case of metals but modern technology relies\noverwhelmingly on semiconductors, and spin injection from ferroma gnetic metals into semiconductors is hampered by\nthe resistivity mismatch [2] at the interface between the metal and the semiconductor, which causes most of the spin\npolarization to be lost at the interface. Another avenue explores t he use of ferromagnetic semiconductors[3] as spin\ninjection devices, but room temperature ferromagnetism in semico nductors remains a long-term goal.\nAs a result of these challenges, purely electrical means of spin gene ration have been sought. The past decade\nhas witnessed an explosion of interest in electrically-induced spin phe nomena in semiconductors and metals, which\nhas accompanied significant experimental and theoretical progre ss. It has been found that the application of an\nelectric field gives rise to a nonequilibrium spin density in the bulk of the s ample, discussed in Refs. [[4]-[12]], and a\nnonequilibrium spin current, discussed in Refs. [[13]-[85]]. The appeara nce of a steady-state spin density in an electric\nfield was predicted several decades ago [4] and subsequently obse rved in tellurium. [5] This work was followed by a\nnumber of theoretical [6, 7, 8, 9] and experimental [10, 11, 12] st udies. Spin currents also have a long history, their\ntheoretical prediction stretching back almost four decades to th e work of Dyakonov.[13] The theoretical work was,\nhowever, not followed by experiments, and the topic was silent for o ver a quarter of a century, until the theory picked\nup again, led by Hirsch [14], Murakami et al.[15] and Sinova et al.[16] During the recent surge ofinterest in electrically-\ninduced spin phenomena the focus has been principally on spin curren ts, with the ultimate goal of measuring spin\nflow in a given direction. In the absence of a spin-ammeter , initially the only way of detecting a spin current relied on\nmeasuring the spin accumulation it produced at the edge of the samp le. This procedure, as will be discussed below,\nis complicated by the fact that the spin current is often not well defi ned and its relationship to spin accumulation\nis not clear. Thus one can detect a spin accumulation at the edge of t he sample as a result of the spin current, but\ninferring the size of the spin current from such a measurement is be yond current understanding. Nevertheless, in just\na few years experiment has been surging ahead. First, in a recent e xperiment, Crooker and Smith [17] imaged the\nflow of spin in semiconductors in the presence of electric and magnet ic fields, as well as strain. Following that, spin\naccumulations resulting from spin currents were measured by seve ral groups. [18, 19, 20, 21, 22, 23, 24] In what is\nsurely a new generation of experiments, Valenzuela and Tinkham[25 ] made use of the fact that a spin current gives\nrise to a transversecharge current. Relying on a nonlocal techniq ue, their group successfully detected a charge current\nflowing as a result of a spin current in Al metal. Cui et al.[26] also observed an electrical current induced by a spin\ncurrent, though in that case the spin current was generated opt ically. This technique was used by several groups2\n[27, 28, 29, 30, 31] shortly afterwards in the study of platinum- an d gold-based structures. In gold an unusually large\nspin current was observed[30] even at room temperature. Room t emperature spin currents were also reported by\nSternet al[21] in ZnSe. Thus, from their initial observation four years ago, sp in currents have begun an experimental\nrevolution, and the field has matured to the point where experiment and theory interact constructively.\nThe phenomena discussed in this review are all related to spin-orbit in teractions, which are present in the band\nstructure and in potentials due to impurity distributions. Band stru cture spin-orbit interactions are frequently the\nmost important factor determining spin dynamics in solids, and are pa rticularly important in semiconductors, where\ncarriersinvolvedin steady-stateprocesseshavewavevectorsw ell within the first Brillouinzone. At these wavevectors,\nwhile the spin-orbit contribution to the carrierenergy is still smaller t han the kinetic energy, it plays an important role\nin determining the energy spectrum. Band structure spin-orbit co upling may arise from the inversion asymmetry of\nthe underlying crystal lattice [86], from the inversion asymmetry of the confining potential in two dimensions [87], and\nmay be present also in inversion symmetric systems. [88] Spin-orbit c oupling is also in principle present in impurity\npotentials. Spin-orbit interactionsin the impurity potentialscauses kewscattering, orasymmetricscatteringofup and\ndown spins, giving scattering-dependent or extrinsic contributions to spin currents. Finallly, spin-orbit interactions\ncause a modification ofthe position and velocity operators, which aff ects the interactionwith an electric field as well as\nthe scattering term. This last mechanism is referred to as side jump and is typically classified as extrinsic although its\ncontributions are manifold. These extrinsic mechanisms were discov ered in the anomalous Hall effect [89, 90, 91, 92]\nbut have acquired a new relevance in recent years in electrically-indu cedspinphenomena.\nIn charge transport, the steady state is characterized by a non equilibrium density matrix that is divergent in the\nclean limit, indicating a competition between the electric field, accelera ting charge carriers, and scattering, which\ninhibits their forward motion. The density matrix is a scalar and the co rrection to it caused by the electric field is\n∝n−1\ni, whereniis the impurity density. When considering the steady state of syste ms with spin-orbit interactions\nthree different situations are distinguished: the case when only ext rinsic mechanisms are important, the case when\nonly band-structure mechanisms are important, and the case whe n both extrinsic and band-structure mechanisms are\nimportant. The differences between these cases are significant an d nontrivial.\nIf band structure spin-orbit interactions are negligible and only ext rinsic mechanisms are important, there is no\nspin precession and the correction to the spin density matrix due to an electric field is very similar to the correction to\nthe scalar density matrix in charge transport. This correction is ∝n−1\niand it gives rise to a steady-state spin current\nbut not to a steady-state spin density. There have been many stu dies of extrinsic spin currents in semiconductors\n(Refs. [[77]-[79]]) and metals (Refs. [[14],[84]].)\nNonequilibrium corrections that arise as a result of band structure spin-orbit coupling in crystal Hamiltonians\nrepresent a different kind of interplay between the electric field and scattering processes. The spin-orbit splitting\nof the bands gives rise to spin-dependent scattering even from sp in-independent potentials, and spin currents and\nspin densities in an electric field arise from linearly independent contrib utions to the density matrix. The steady-\nstate density matrix contains a contribution due to precessing spin s and one due to conserved spins. Steady-state\ncorrections ∝n−1\niare associated with the absenceof spin precession and steady-state corrections independent of ni\nare associated with spin precession. Steady-state corrections ∝n−1\nigive rise to spin densities in external fields while\nsteady-state corrections independent of nigive rise to spin currents in external fields. Scattering between th ese two\ndistributions induces significant corrections to steady-state spin currents, and may cause spin currents to vanish.\nThe phrase spin current refers to the flow of spins across a sample and, if spin were conserv ed, one could distinguish\nbetween spin-up and spin-down charge currents. In systems with band structure spin-orbit interactions the spin\ncurrent is not well defined [32, 33, 34] and its relationship to spin acc umulation is unclear. Spin transport in these\nsystems usually does not involve charge transport as the charge c urrents in the direction of spin flow cancel out.\nIn addition, since spin currents may be accompanied by steady-sta te spin densities, this makes it more difficult to\nseparate experimentally the signals due to these contributions. Th e study of spin currents in systems with band\nstructure spin-orbit interactions has thus encountered a numbe r of profound physical issues on which no consensus\nexists at present. There has been considerable research has on d efinitions of spin currents in such systems (Refs. [[32]-\n[35]]), the role played by the symmetry of the underlying crystal latt ice,[36] whether spin currents as a result of\nband structure spin-orbit coupling are transport or background currents, [37] the relationship between band structure\nspin currents and spin accumulation, (Refs. [[38]-[45]]) and the form o f the Maxwell equations in systems with band\nstructure spin-orbit coupling.[46, 47] Most theoretical studies ha ve focused on metals, common semiconductors and\nasymmetric quantum wells with band structure spin-orbit interactio ns, such as Refs. [[16] -[76]].\nWhen both band structure and extrinsic mechanisms are present, the situation becomes considerably more compli-\ncated. There is no simple association between spin conservation and spin densities, or between spin precession and\nspin currents. Both spin densities and currents may contain terms ∝n−1\niand independent of ni. A careful analysis\nshows that the presence of band structure spin-orbit interactio ns causes skew scattering and side jump to behave very\ndifferently in the steady state, and under certain circumstances, the contributions of these extrinsic mechanisms to the\nspin current vanish. The interplay of band structure and extrinsic mechanisms has been considered in Refs. [[80]-[83]].3\nThis review will cover steady-state spin densities and currents with in the framework of a density matrix theory.\nSuch a framework is important in order to demonstrate the unity be hind all observed phenomena. I will consider\nlarge, uniform systems, and work in momentum space. Although man y observations in this entry are general, the\ndiscussion will focus on non-interacting spin-1/2electron systems , which are pedagogicallyeasier. This review is based\non a kinetic-equation formalism equivalent to linear response theory and correspondences with other methods will be\nidentified, such as linear response theories based on Green’s funct ions and semiclassical wave packet dynamics.\nThisreviewarticleisstructuredasfollows. Section2will introducead ensitymatrixpictureofspindynamics, briefly\nsketching the way different theories stem from the quantum Liouville equation. In section 3 a brief introduction to\nthe spin-orbit interaction in spin-1/2 systems is given, after which s pin densities and spin currents as a result of band\nstructure spin-orbit interactions are discussed, as well as their r elationship to spin precession. Section 4 is devoted\nto spin densities and spin currents that arise as a result of extrinsic mechanisms, together with the case when both\nextrinsic and band-structure mechanisms are present. Section 5 discusses briefly various definitions of spin currents\nand the relationship between spin currents and spin acumulation. Se ction 6 outlines the role of the underlying crystal\nlattice in the establishment of spin densities and currents and sectio n 5 is concerned with open issues such as the\ndefinition and nature of spin currents, as well as the relationship be tween spin currents and spin accumulation. In\nsection 7 the experimental situation is summarized and in closing futu re directions are discussed. Related topics such\nas the quantum spin-Hall effect [93] and the spin-Hall insulator [94] a re beyond the scope of this review.\nII. DENSITY MATRIX PICTURE OF SPIN DYNAMICS\nElectrically-induced spin phenomena encompass a wide variety of pro cesses that have been studied using different\nmethods. In order to bring out the unity behind the processes invo lved, as well as behind the theoretical approaches\nemployed, it will be useful to have a unified frameworkin which electric ally-induced spin phenomena can be discussed.\nIn this section such a framework will be constructed beginning with t he quantum Liouville equation. This equation,\nin one form or another, is the starting point of most theories of spin dynamics, and this section will outline the way\nvarious approaches are related to each other. The focus will be on systems with large homogeneous systems with long\nmean free paths, and diffusion will not be considered explicitly. Electr ic fields will be assumed uniform.\nA. Quantum Liouville equation\nA system of non-interacting spin-1/2 electrons is represented by a one-particle density operator ˆ ρ. The expectation\nvalue of an observable represented by a Hermitian operator ˆOis given by Tr(ˆ ρˆO), where Tr denotes the most general\noperator trace, which involves summation over discrete degrees o f freedom and integration over continuous ones. The\nusual matrix trace will be denoted by tr. The dynamics of ˆ ρare described by the quantum Liouville equation,\ndˆρ\ndt+i\n/planckover2pi1[ˆH+ˆUdis+eE·ˆr,ˆρ] = 0. (1)\nThe Hamiltonian ˆHcontains contributions due to the kinetic energy and spin-orbit cou pling, while ˆris the position\noperator. The effect of the lattice-periodic potential of the ions is taken into account through a replacement of\nthe carrier mass by the effective mass. The potential ˆUdisaccounts for scattering processes, which may be due to\nimpurities, phonons, surface roughness, or other perturbation s. This review focuses on impurity scattering, as the\neffects discussed are frequently observed at low temperatures, where scattering due to phonons may be neglected. It\nis assumed that the electric field is small and a solution is sought to firs t order in the electric field. The solution of the\nLiouville equation (1) in the absence of the external field is provided b y ˆρ0, the Fermi-Dirac function. The correction\ndue to the external field, ˆ ρE, satisfies\ndˆρE\ndt+i\n/planckover2pi1[ˆH+ˆUdis,ˆρE] =−i\n/planckover2pi1[eE·ˆr,ˆρ0]. (2)\nEquation(2)isprojectedontoacompletesetoftime-independent statesofdefinitewavevector {|ks∝an}bracketri}ht}. Thesestatesare\nnot assumed to be eigenstates of ˆH. The matrix elements of ˆ ρin this basis will be written as ρkk′≡ρss′\nkk′=∝an}bracketle{tks|ˆρ|k′s′∝an}bracketri}ht,\nwith corresponding notations for the matrix elements of ˆHandˆUdis. Spin indices will not be shown explicitly in the\nsubsequent discussion, and it will be understood that the quantitie sρkk′,Hkk′, andUdis\nkk′are matrices in spin space.\nρkk′is referred to as the density matrix. With our choice of basis functio ns of definite wave vector, matrix elements of\nthe Hamiltonian Hkk′=Hkδkk′arediagonalin k. However,sincethe Hamiltoniancontainsspin-orbitcouplingterms,\nmatrix elements Hkare generally off-diagonal in spin space. Matrix elements of the scat tering potential Udis\nkk′are4\noff-diagonal in k. Matrix elements diagonal in kin the scattering potential would lead to a redefinition of Hk, which\nis analogous, in Green’s function formalisms, to the offset introduce d by the real part of the self energy. Scattering\nis assumed elastic and impurities uncorrelated, and the normalization is such that the configurational average of\n∝an}bracketle{tks|ˆUdis|k′s′∝an}bracketri}ht∝an}bracketle{tk′s′|ˆUdis|ks∝an}bracketri}htis (ni|Ukk′|2δss′)/V, whereniis the impurity density, Vthe crystal volume and Ukk′the\nmatrix element of the potential of a single impurity. Configurational averages over terms of higher order in ˆUdisare\nperformed in a similar fashion. [96] It is assumed that εFτ//planckover2pi1≫1, where εFis the Fermi energy and τa characteristic\nscattering time. This is equivalent to the assumption that the carrie r mean free path is much larger that the de\nBroglie wavelength. None of the methods presented below are valid o nceεFτp//planckover2pi1becomes comparable to unity.\nSpin densities and spin currents are obtained as expectation values of spin and spin current operators. The spin\noperator is given by sσ= (/planckover2pi1/2)σσ, whereσσis a Pauli spin matrix. The spin current operator will be taken to be\nˆJσ\ni= (1/2){sσ,vi}, where the velocity operator is vi= (1//planckover2pi1)∂Hk/∂ki. Both operators are diagonal in k. One is\ntherefore primarily interested in the part of the density matrix diag onal ink. With this observation in mind, ρkk′is\ndivided into a part diagonal in kand a part off-diagonal in kasρkk′=fkδkk′+gkk′, where, in gkk′, it is understood\nthatk∝ne}ationslash=k′. Further, fkis decomposed into a scalar part and a spin-dependent part, fk=nk11 +Sk, with11 the\nidentity matrix. In an electric field, fk=f0k+fEk, wherefEkis a correction linear in Eand has a corresponding\ndecomposition into a scalar part and a spin-dependent part. At the end one takes the trace of the spin and spin\ncurrent operators with fEkgiven by Eq. (4). The final result is usually expressed in many ways.\nIt should be noted that the electric field induces coherence betwee n bands. In equilibrium the Fermi-Dirac function\nf0kcommutes with the Hamiltonian and is stationary. However the sourc e term due to Ein Eq. (2) does not in\ngeneral commute with the Hamiltonian. If the problem is considered in the basis of eigenstates of Hk, the source\nterm couples different energy bands. If a basis is used in which one sp in component is a good quantum number, then\nthe source term couples up and down spins.\nB. Linear response Green’s function formalism\nThe linear response Green’s function formalism is closely related to th e density matrix formalism but relies on a\nsomewhat different way of regarding the problem [96]. The electric fi eld-induced correction to the density matrix ˆ ρE\nis found by applying the time evolution operator corresponding to th etotalHamiltonian ˆH+ˆUdis, incuding disorder,\nin Eq. (2). The matrix elements of the time evolution operator in the b asis{|ks∝an}bracketri}ht}constitute the Green’s function\nGss′\nkk′(t), defined by\nGss′\nkk′(t) =∝an}bracketle{tks|e−i(ˆH+ˆUdis)t//planckover2pi1|k′s′∝an}bracketri}ht. (3)\nSpin indices are again suppressed. All the information about the sys tem is contained in the Green’s function, which\nis also referred to as the propagator, since it gives the amplitude fo r a particle to propagate in time tfrom state\n|ks∝an}bracketri}htto state|k′s′∝an}bracketri}ht. The Green’s function is also the kernel of the Schrodinger equatio n and represents the response\nof the system to a perturbation that is localized in real space or mom entum space. The total perturbation is built\nup as a sum of localized ones. The problem can be formulated in any bas is but for our purposes it is still the most\nconvenient to work in the basis of eigenstates of Hsok. In general the Green’s function is a matrix in spin space. For\npractical purposes one also defines the retarded and advanced G reen’s functions, GRandGA, asGR=−iGθ(t) and\nGA=iGθ(−t), withθthe Heaviside step function. The electric field-induced correction t o the part of the density\nmatrix diagonal in kis\nfEk=−ieE\n/planckover2pi1·/summationdisplay\nk′/integraldisplay∞\n−∞dt′GR\nkk′(t′)[ˆr,ˆρ0]k′GA\nk′k(−t′) (4)\nAt this stage it is easiest to carry out a Fourier transformation with respect to time. The integration in Eq. (4) has\nthe same form except the time variable is replaced by the frequency variable ω\nfEk=−ieE\n/planckover2pi1·/summationdisplay\nk′/integraldisplay∞\n−∞dωGR\nkk′(ω)[ˆr,ˆρ0]k′GA\nk′k(−ω). (5)\nThe correction fEkto the density matrix needs to be averaged over impurities, and it is a ssumed henceforth that\nonly the impurity averaged density matrix is of interest. We denote t his impurity average by the symbol ∝an}bracketle{t∝an}bracketri}ht, but the\nimpurity-averaged density matrix will be abbreviated as ∝an}bracketle{tfEk∝an}bracketri}ht ≡fEk. The full Green’s function is determined by the\ntotal Hamiltonian ˆH+ˆUdisincluding the scattering potential. The Green’s function is expanded in the scattering5\npotential, and one is only interested in the Green’s function averaged over impurity configurations. For uncorrelated\nimpurities the impurity-averaged Green’s function, denoted by ∝an}bracketle{tG∝an}bracketri}ht, obeys the recursion relation\n∝an}bracketle{tGR/A(k,ω)∝an}bracketri}ht=GR/A\n0(k,ω)+GR/A\n0(k,ω)ΣR/A(k,ω)GR/A(k,ω) (6)\nin which G0is the equilibrium Green’s function, corresponding to the Hamiltonian ˆHwithout disorder, and the self\nenergy Σ is typically evaluated in the first Born approximation\nΣR/A(k,ω) =ni/integraldisplayddk′\n(2π)d|Ukk′|2GR/A\n0(k′,ω) (7)\nwithdthe dimensionality of the system. The correction fEkto the density matrix depends on the product of two\nGreen’s functions. After the average over impurity configuration s we obtain terms which are expressible as a product\nof two individual impurity-averagedGreen’s functions and cross te rms, which connect different Green’s functions, and\nwhich are not expressible as a product. These form a vertex Γ kk′defined by\n∝an}bracketle{tGR(k,ω)GA(k′,ω)∝an}bracketri}ht=∝an}bracketle{tGR(k,ω)∝an}bracketri}ht∝an}bracketle{tGA(k,ω)∝an}bracketri}ht+Γkk′(ω)∝an}bracketle{tGR(k′,ω)∝an}bracketri}ht∝an}bracketle{tGA(k′,ω)∝an}bracketri}ht. (8)\nThe vertex is in general expressed as a sum of different classes of d iagrams, of which the most relevant to transport\nare the ladder diagrams and the maximally crossed diagrams. It also s atisfies the recursion relation\nΓkk′(ω) =Mkk′(ω)+/integraldisplayddk′′\n(2π)dMkk′′(ω)∝an}bracketle{tGR(k′′,ω)∝an}bracketri}ht∝an}bracketle{tGA(k′′,ω)Γk′′k′(ω)∝an}bracketri}ht (9)\nwhereMkk′(ω) can alsobe expressed in terms ofdiagramsas shown in Ch. 8 ofRef. [[96]]. Once the impurity-averaged\nGreen’sfunction ∝an}bracketle{tG∝an}bracketri}htandthe vertexareknown, fEkcan be found. It will be noticed that in the linearresponseGreen’s\nfunction approach disorder is contained in the self energy Σ( k) and in the vertex Γ kk′. The response function contains\nall the disorder up to the desired approximation. Once the Green’s f unction and the vertex function are found, the\ntrace with the spin and spin current operators is taken.\nC. Kinetic equation formalism\nThe kinetic equation formalism can be derived from the quantum Liouv ille equation, as will be done below, or using\na Keldysh Green’s function formalism.[57] The kinetic equation obtaine d is naturally the same. Starting from the\nquantum Liouville equation, this can be broken down into equations fo rfkandgkk′\ndfEk\ndt+i\n/planckover2pi1[Hk,fEk] =−i\n/planckover2pi1[ˆU,ˆgE]kk−i\n/planckover2pi1[eE·ˆr,ˆρ0]kk, (10a)\ndgEkk′\ndt+i\n/planckover2pi1[ˆH,ˆgE]kk′=−i\n/planckover2pi1[ˆU,ˆfE+ ˆgE]kk′. (10b)\nIn the first Born approximation the solution to Eq. (10b) can be writ ten as\ngEkk′=−i\n/planckover2pi1/integraldisplay∞\n0dt′e−iˆHt′//planckover2pi1/bracketleftBig\nˆU,ˆfE(t−t′)/bracketrightBig\neiˆHt′//planckover2pi1|kk′. (11)\nSinceεFτp//planckover2pi1≫1, we shall expand ˆf(t−t′) in the time integral around tand, noting that terms beyond ˆf(t) are of\nhigher order in the scattering potential, we shall only retain the firs t term,ˆf(t). The equation for fkthen becomes\ndfEk\ndt+i\n/planckover2pi1[Hk,fEk]+ˆJ(fEk) =eE\n/planckover2pi1·/parenleftbigg∂f0k\n∂k−i[R,f0k]/parenrightbigg\n(12a)\nin which the scattering term ˆJ(fk) is given by\nˆJ(fEk) =1\n/planckover2pi12/integraldisplay∞\n0dt′/bracketleftBig\nˆU,e−iˆHt′//planckover2pi1/bracketleftBig\nˆU,ˆfE(t)/bracketrightBig\neiˆHt′//planckover2pi1/bracketrightBig\nkk(12b)\nand the source term on the RHS contains the covariant derivative with respect to k, which takes into account the\nfact that the basis functions themselves may depend on k. It includes the gauge connection matrix R≡Rss′=6\n∝an}bracketle{tks|i∂/∂k|ks′∝an}bracketri}ht. Theresponsefunction, aswellasthescatteringtermneedtobe averagedoverimpurityconfigurations,\nand the notation fEkis used henceforth as above to denote the impurity-averaged den sity matrix. The scattering\nterm can be expanded further in the basis in spin space spanned by s pin eigenstates | ↑∝an}bracketri}htand| ↓∝an}bracketri}ht(the Pauli basis.)\nIn this basis the connection matrix Rvanishes and the scattering potential Ukk′=Ukk′11 is diagonal in spin space.\nConverting sums over wave vector into integrals/summationtext\nk′→V/integraltextddk′\n(2π)d, and performing the time integral the scattering\nterm can be expressed in the form ( ˆJ0+ˆJb)(nk)+ˆJ0(Sk), with\nˆJ0(Sk) =2πni\n/planckover2pi1/integraldisplayddk′\n(2π)d|Ukk′|2(Sk−Sk′)δ(ε0k−ε0k′) (13a)\nˆJb(nk) =2πni\n/planckover2pi1/integraldisplayddk′\n(2π)d|Ukk′|2(nk−nk′)1\n2σ·(Ωk−Ωk′)∂\n∂ε0kδ(ε0k−ε0k′).\nThe term ˆJb(nk) in Eq. (13b) illustrates the fact that, when spin-orbit interaction s are present in the band structure,\neven spin-independent scattering potentials give rise to spin-depe ndent terms in the scattering integral. For potentials\ndiagonalin spinspaceandspin-degeneratebands, Eq.(12b) simplifi estothe customaryFermi’sgoldenrule. Therefore,\nEq. (12) is a generalization of Fermi’s golden rule that explicitly takes in to account the spin degree of freedom.\nD. Boltzmann-wave packet formalism\nThe density matrix is the most complete description of a physical sys tem. It contains all the relevant physics,\nincluding interband coherence due to the electric field and scatterin g. It represents the system as a whole and\naccounts for the individual particle motion, which occurs between c ollisions, as well as for the distribution of electrons\nin phase space and the way it is altered by collisions. As was shown abov e, linear response theories based on Green’s\nfunctions and theories based on kinetic equations originate directly from the Liouville equation for the density matrix.\nA third approach, which may be called a semiclassical or Boltzmann-wa ve packet approach, separates the motion\nof the charge carriers from their distribution in phase space. To de termine the dynamics of the charge carriers one\nenvisages them as being represented by wave packets, and the ph ase space distribution of wave packets is given by a\nfunction of the Boltzmann form. One use of the word semiclassical , which will be adopted in this work, is to refer to\ntheories which consider the position and momentum of a particle simult aneously. Semiclassical approaches exploit the\nsmooth variation of transport fields on atomic length scales in order to provide intuitive descriptions of steady-state\nprocesses, which can be easily extended to cover inhomogeneous s ystems and spatially dependent fields.\nThe semiclassical approach is also closely related to the density matr ix and the Liouville equation. Firstly, in order\nto separate the particle dynamics and the phase space distribution one must be able to label the particles by a band\nindex, therefore the theory must be formulated in the basis of eige nstates of Hk. Since these eigenstates are usually\nk-dependent it is the covariant derivative that enters the source t erm in the kinetic equation. The point at which\nsemiclassical theory branches out of the kinetic equation is Eq. (12 a). In order to turn the density matrix into a\nBoltzmann distribution function one takes the diagonal elements of this equation. If we consider for simplicity a spin-\n1/2system, which ismade up oftwobandsthat will be called 1and2, we canlabel the diagonalelements ofthe density\nmatrix as f1kandf2k. These become the Boltzmann distributions for bands 1 and 2. The c ommutator [ Hk,fk] has\nno diagonal elements and ˆJ(fk) reduces to one scattering term for each band, which is given by Fe rmi’s Golden Rule,\nplus interband scattering terms which couple f1kandf2k. In this way Eq. (12a) reduces to two coupled Boltzmann\nequations for the two bands and the phase space distribution has b een separated out of the kinetic equation.\nIn reducing the kinetic equation to a series of Boltzmann equations o ne neglects interband coherence due to the\nelectric field, and this must be recovered elsewhere. In the semiclas sical approach, this occurs in several ways. Firstly,\none needs to consider the nature of the carriers in band sand invoke explicitly the fact that carriers are described\nby wave packets. This is the point at which the formalism turns into a t rue semiclassical theory. The construction\nof a wave-packet representing a charge and spin carrier in band s, which has real and k-space coordinates ( rcs,kcs),\nhas been thoroughly treated by Sundaram and Niu [97] and its exten sion to multiple bands was carried out in Ref.\n[[98]]. The macroscopic distribution associated with a quantum mechan ical operator ˆOis no longer given simply by\nthe expectation value of the operator ˆO, but involves explicitly a sum over wave packets centered at ( rcs,kcs). For\nexample the spin density distribution is\nSσ(r,t) =/summationdisplay\ns/integraldisplay\nd3kcs/integraldisplay\nd3rcsfs(kcs,t)∝an}bracketle{tδ(r−ˆ r)ˆsσ∝an}bracketri}hts, (14)\nwhere the bracket indicates quantum mechanical average over th e wavepacket with charge centroid ( rcs,kcs). An\nanalogous expression exists for the spin current distribution. The wave-packet expectation values ∝an}bracketle{t∝an}bracketri}htsin Eq. (14)7\nr rCharge Spin\nc s\nFIG. 1: For a particle of finite extent the charge and spin dist ributions in real space are in general do not coincide. The sa me\nis true of the charge and spin distributions in reciprocal sp ace.\neventually yield functions of ( rcs,kcs), the dynamics of which are discussed below. It may appear contra dictory that\none has to integrate over the real-space coordinates of the wave packetsrcseven in the case of homogeneous systems\nstudied in this work, where the distribution function is not a function ofrcs. This is because even in a homogeneous\nsystem of the center of charge and the center of spin are not the same, so variations of the spin distribution on the\nspatial scales comparable to that of the wave packet need to be ta ken into account, as shown in Fig. 1. All these\nterms must be considered in order to reach agreement with approa ches based directly on the Liouville equation.\nIn a constant uniform electric field Ethe coordinates of the wave packet center ( rcs,kcs) drift according to the\nsemiclassical equations of motion [97]\n/planckover2pi1˙kcs=−eE\n/planckover2pi1˙rcs=∂εs\n∂kc+eE×Fs,(15)\nwhereFs=∇k×RssrepresentstheBerry,orgeometricalcurvature[97], with ∇kthegradientoperatorinmomentum\nspace. A careful analysis reveals that the Berry curvature also r epresents coherence between bands and must be taken\ninto account in order to obtain agreement with approaches based d irectly on the Liouville equation. It is emphasized\nthat the Berry curvature is not a result of the particles being desc ribed by wave packets, which have a finite extent in\nreal and momentum space. Rather, it emerges from the phase of t he Bloch wave functions involved in the construction\nof the wave packets, implying that this formulation of semiclassical t ransport theory is a useful tool for capturing\nBerry-phase effects.\nThe electric field and disorder potential also give rise to a nonadiabat ic mixing of the bands such that the basis\nstates|ks∝an}bracketri}htbecome|˜ks∝an}bracketri}ht, given by\n|˜ks∝an}bracketri}ht=|ks∝an}bracketri}ht+/summationdisplay\nk′,s′/negationslash=s[eEδkk′−(∇Udis)kk′]·Rs′s\nεks−εks′|ks′∝an}bracketri}ht, (16)\nwhere the |ks∝an}bracketri}htare the unperturbed eigenstates. The |˜ks∝an}bracketri}htalso form a complete set. The distribution functions\nfs(kcs,t) are made up of an equilibrium part and a part linear in the electric field, fs(kcs,t) =f0s(kcs,t)+fEs(kcs,t),\nand the expectation values ∝an}bracketle{t∝an}bracketri}htsin Eq. (14) also have terms of zeroth and linear orders in the electric field. Therefore\nto first order in the field the expectation value of operator ˆOcontains terms arising from the product of fEs(kcs,t)\nwith the zeroth order term in ∝an}bracketle{t∝an}bracketri}htsas well as from the equilibrium f0s(kcs,t) multiplied by the term in ∝an}bracketle{t∝an}bracketri}htslinear in the\nelectric field.\nThe above exposition shows that interband coherence lost in passin g from the density matrix to the Boltzmann\napproach is thus recovered in many ways. It is present in the Berry curvature, in the nonadiabatic correction to\nthe wave functions, and in the interband scattering terms coupling the two distribution functions. The construction\nof a correct semiclassical theory that contains all the physics rele vant to transport is seen to entail a sizable effort,\nparticularly in the case of transport of non-conserved quantities such as spin.8\nIII. BAND-STRUCTURE SPIN DENSITIES AND CURRENTS\nThis work will now focus on the kinetic equation formalism and determin e the spin densities and spin currents\ninduced by external electric fields under a variety of circumstance s. It is most straightforward to begin with effects\noriginating in the band structure. The Hamiltonian of spin-1/2 electr on systems typically contains a kinetic energy\nterm and a spin-orbit coupling term, Hk=/planckover2pi12k2\n2m∗+Hso\nk, wherem∗is the electron effective mass. In spin-1/2 electron\nsystems, band structure spin-orbit coupling can always be repres ented as a Zeeman-like interaction of the spin with\na wave vector-dependent effective magnetic field Ωk, thusHso\nk= (/planckover2pi1/2)σ·Ωk. Common examples of effective fields\nare the Rashba spin-orbit interaction, [87] which is often dominant in quantum wells with inversion asymmetry, and\nthe Dresselhaus spin-orbit interaction, [86] which is due to bulk inver sion asymmetry.\nAn electron spin at wave vector kprecesses about the effective field Ωkwith frequency Ω k//planckover2pi1≡ |Ωk|//planckover2pi1and is\nscattered to a different wavevectorwithin a characteristicmomen tum scatteringtime τ. Within the range εFτ//planckover2pi1≫1,\nthe relative magnitude of the spin precession frequency Ω kand inverse scattering time 1 /τdefine three qualitatively\ndifferent regimes. In the ballistic or clean regime no scattering occur s and the temperature tends to absolute zero,\nso thatεFτ//planckover2pi1→ ∞and Ω kτ//planckover2pi1→ ∞. The weak scattering regime is characterized by fast spin precess ion and little\nmomentum scattering due to, e.g., a slight increase in temperature, yielding εFτ//planckover2pi1≫Ωkτ//planckover2pi1≫1. In the strong\nmomentum scattering regime εFτ//planckover2pi1≫1≫Ωkτ//planckover2pi1.\nThe kinetic equation (12a) has a scalar part which determines nEk\n∂nEk\n∂t+ˆJ0(nEk) =eE\n/planckover2pi1·∂n0k\n∂k. (17)\nThe solution of this equation is given by the well-known expression\nnEk=eEτp\n/planckover2pi1·∂n0k\n∂k, (18)\nin other words, nEkdescribes the shift of the Fermi sphere in the presence of the elec tric field E. The expression\nfor the momentum relaxation time τpis a little different depending on the dimensionality of the system. Using γto\ndenote the relative angle between kandk′,\n1\nτp=nimk\n2π/planckover2pi13/integraldisplayπ\n0dγ|Ukk′|2sinγ(1−cosγ) (3D) (19a)\nnim\n2π/planckover2pi13/integraldisplay2π\n0dγ|Ukk′|2(1−cosγ) (2 D). (19b)\nThe spin-dependent part of the nonequilibrium correction to the de nsity matrix SEkis the spin density induced by\nE. Its time evolution is governed by SEkis\n∂SEk\n∂t+i\n/planckover2pi1[Hk,SEk]+ˆJ0(SEk) =eE\n/planckover2pi1·∂S0k\n∂k−ˆJb(nEk). (20)\nSpin-dependent scattering gives rise to a renormalization of the dr iving term in the equation for SEkwith no analog\nin charge transport, see Eq. (17).\nSince an electron spin at wave vector kprecesses about Ωk, the spin can be resolved into components parallel and\nperpendicular to Ωk. In the course of spin precession the component of the spin paralle l toΩkis conserved, while the\nperpendicular component is continually changing. It will prove usefu l in our analysis to divide the spin distribution\ninto a part representing conserved spin and a part representing p recessing spin. The effective source term, which\nenters the RHS of Eq. (20), is divided into two parts, ( eE//planckover2pi1)·∂S0k/∂k−ˆJb(nEk) = ΣEk/bardbl+ΣEk⊥. ΣEk/bardblcommutes\nwithHso\nkwhile Σ Ek⊥isorthogonal to it and tr(Σ Ek⊥Hso\nk) = 0. Projections onto and orthogonal to Hso\nkare most\neasily carried out by defining projectors P/bardblandP⊥by their actions on the basis matrices σias described in Ref. [[36]].\nSEkis likewise divided into two terms: SEk/bardbl, commuting with the spin-orbit Hamiltonian and SEk⊥, orthogonal to\nit.SEk/bardblis the distribution of conserved spins while SEk⊥is the distribution of precessing spins. Equation (20) is\ndivided into separate equations for SEk/bardblandSEk⊥\n∂SEk/bardbl\n∂t+P/bardblˆJ0(SEk) = ΣEk/bardbl, (21a)\n∂SEk⊥\n∂t+i\n/planckover2pi1[Hk,SEk⊥]+P⊥ˆJ0(SEk) = ΣEk⊥. (21b)9\nThe absence of the commutator [ Hk,SEk/bardbl] = 0 in Eq. (21a) indicates the absence of spin precession, while the\ncommutator [ Hk,SEk⊥] in Eq. (21b) represents spin precession. In order to solve Eqs. ( 21a) and (21b) for arbitrary\nscattering, it is necessaryto expand SEk/bardblandSEk⊥inni, asSEk/bardbl=S(−1)\nEk/bardbl+S(0)\nEk/bardbl+O(ni) andSEk⊥=S(0)\nEk⊥+O(ni).\nThis is an expansion in the parameter /planckover2pi1/(Ωkτp) and is most suited to systems in the weak scattering regime. The\nexpansion of SEk/bardblstarts at order −1, a fact which can be understood by inspecting Eq. (21a). In the s teady state\nthe time derivative drops out, and the operator ˆJ0is first order in ni, while the right-hand side is independent of ni.\nAs a result, the expansion of the solution must start at order −1. Equation (21b) for SEk⊥tells us that, since Hkis\nindependent of niand the right hand side is also independent of ni, the expansion of SEk⊥must start at order zero.\nEquation (21a) can be solved iteratively for any scattering\nS(−1)\nEk/bardbl= ΣEk/bardblτ0+P/bardbl/parenleftbiggm∗\n2π/planckover2pi13/integraldisplay\ndθ′|Ukk′|2ΣEk/bardbl/parenrightbigg\nτ2\n0+... (22)\nwhereτ0=nim∗/integraltext\ndθ′|Ukk′|2/(2π/planckover2pi13) is the quantum lifetime of the carriers. Above and henceforth inte grals over\nwave vectors will be represented as two-dimensional, and θ′will refer to the polar angle of k′. The extension to three\ndimensions is straightforward. The equations for higher orders in niare easily deduced. However, the term of order\n−1 is by far the dominant one in the weak momentum scattering regime a nd is expected to be dominant over a wide\nrange of strengths of the scattering potential.\nIt is evident that the steady state for conserved spins involves no spin precession, and that the correction SEk/bardbl\ndepends explicitly on the nonequilibrium shift in the Fermi surface. In addition, scattering terms contain only the\neven function |Ukk′|2. As a result, SEk/bardbldoes not give rise to a spin current. Inspection of Eq. (22) shows t hat integrals\nof the form/integraltext\ndθˆJσ\niSEk/bardblcontain an odd number of powers of kand are therefore zero. In the absence of impurity\nspin-orbit interactions, the distribution of conserved spins can giv e no spin current. It can, however, give rise to a\nnonequilibrium spin density since integrals of the form/integraltext\ndθˆsσSEk/bardblcontain an even number of powers of kand may\nbe nonzero.\nThe leading term S(0)\nEk⊥is found to be\nS(0)\nEk⊥=/planckover2pi1\n2σ·ˆΩk×[ΣEk⊥−P⊥ˆJ0(SEk/bardbl)]\nΩk, (23)\nwhere we have written Σ Ek⊥= (1/2)ΣEk⊥·σandSEk/bardbl= (1/2)SEk/bardbl·σ. The result expressed by Eq. (23) is valid\nfor any elastic scattering. An argument similar to that given above s hows that S(0)\nEk⊥cannot lead to a nonequilibrium\nspin density (although, as will be shown below, higher-order terms in SEk⊥can contribute to the spin density). For,\ntaking the expectation value of the spin operator, one arrives at in tegrals of the form/integraltext\ndθˆsσS(0)\nEk⊥, which involve odd\nnumbers of powers of kand are therefore zero. This term in the distribution of precessing spin does, however, give\nrise to nonzero spin currents, since integrals if the form/integraltext\ndθˆJσ\niSEk⊥contain an even numbers of powers of kand\nmay be nonzero. Consequently, in the absence of spin-orbit couplin g in the scattering potential, nonequilibrium spin\ncurrents arise from spin precession.\nThe dominant contribution to the nonequilibrium spin density in an elect ric field exists because in the course of spin\nprecession a component of each individual spin is preserved. For an electron with wave vector k, this spin component\nis parallel to Ωk. In equilibrium the average of these conserved components is zero . However, when an electric field is\napplied, the Fermi surface is shifted, and the average of the cons erved spin components may be nonzero, as illustrated\nin Fig. 2. This intuitive physicalargumentexplains why the nonequilibriu m spin density ∝τ−1\npandrequiresscattering\nto balance the drift of the Fermi surface. It is interesting to note , also, that, although spin densities in electric fields\nrequire the presence of band structure spin-orbit interactions a nd therefore spin precession, the dominant contribution\narises as a result of the absence of spin precession. Band structu re spin currents on the other hand are associated\nwith displacement of spins. The relation between spin currents and s pin precession was made explicit by Sinova et al.\n[16] Both SEk/bardblandSEk⊥are invariant under time-reversal. As a result, the tensor charac terizing the response of\nspin currents to electric fields is invariant under time reversal, wher eas the tensor characterizing the response of spin\ndensities to electric fields changes sign under time reversal.\nIt will be noticed that the driving term in the equation for the nonequ ilibrium spin distribution SEkis renormalized\nby the term ˆJs(nEk), which accounts for spin-dependent scattering. In addition, Eq . (21b) shows that scattering\nmixes the distributions of conserved and precessing spins. When on e spin at wave vector kand precessing about\nΩkis scattered to wave vector k′and precesses about Ωk′, its conserved component changes, a process which\nalters the distributions of conserved and precessing spin. Conseq uently, scattering processes in systems with spin-\norbit interactions cause a renormalization of the driving term for th e spin distribution as well as scattering between\nthe conserved and precessing spin distributions. Contributions du e to these two processes are contained in vertex10\nkxky\nkxky\n0 00 0E=0 E>0 (b) (a)\nFIG. 2: Effective field Ωkat the Fermi energy in the Rashba model (a) without ( E= 0) and (b) with ( E >0) an external\nelectric field.\ncorrections to spin-dependent quantities found in Green’s functio ns formalisms. For short-range impurities it is easy\nto show [36, 75] that the scattering correction P⊥ˆJ0(SEk/bardbl) to Eq. (23) depends on the steady state spin density, and\nit becomes evident that the existence of a nonzero nonequilibrium sp in density does affect the spin current. This fact\nwas first pointed out by Raimondi [75] for the Rashba model.\nThe spin-Hall current was initially determined for spin-3/2 holes in GaA s [15] and for an asymmetric quantum well\n[16] in which the spin-orbit interaction is described by the Rashba mod el. The latter calculation yielded a spin-Hall\nconductivity σz\nxy=e/(8π), in which σi\njkis understood as referring to spin component iflowing in direction jin\nresponse to an electric field applied along k. However, it was subsequently shown that a more careful treatm ent of\ndisorder renormalizes this result to zero. In fact in two dimensions, for Hamiltonians linear in wave vector, the spin\ncurrent vanishes for short-range impurities (Refs. [[50]-[55]]), for small-angle scattering [56, 57] and in general for any\nelastic scattering. [33] However, it does not vanish in a generalized R ashba model as was shown by Krotkov and Das\nSarma. [68] For spin-orbit Hamiltonians characterized solely by one a ngular Fourier component Nthe spin current\n∝N. [57]\nAb initio calculation of band structure spin currents were also perfo rmed by Guo et al.[72] in semiconductors and\nby Yaoet al.[73] in semiconductors and simple metals such as tungsten, platinum a nd gold. Spin transport in metals\nusually requires complex band structure calculations, which are typ ically done numerically. Yao et al.[73] found that\nthe band structure spin-Hall effect in metals can be significantly larg er than in semiconductors and that the spin-Hall\nconductivity can even undergo sign changes under certain circums tances.\nA. Spin adiabatically following a magnetic field\nThe existence of a spin current in an electric field and its association w ith spin precession, can be understood from\na straightforward argument due to Sinova.[16] A spin precesses ab out an effective magnetic field Ωkwhich depends\nonk. The electric field changes the wave vector kand in the process changes Ωk, so the spin is now precessing\nabout magnetic field that is changing adiabatically. In the adiabatic limit the spin follows the magnetic field. [99] In\nparticular, in 2D if the magnetic field is in the plane of the spin, the spin n ever acquires a significant out-of plane\ncomponent. However the spin in general acquires a small out-of-p lane component. Consider a generalized magnetic\nfieldΩand take the simple example in which the magnetic field ∝bardblˆxand has a small y-component which increases\nlinearly with time, such that Ω y=ǫt. The spin starts out along x, initially parallel to the field. In both cases, sxwill\nbe considered large compared to sy,szand in this section /planckover2pi1= 1. The adiabatic condition means that the magnitude\nof the magnetic field Ω changes little over one revolution of the electr on spin about it. This means that, if we start\natt= 0 and consider a small time increment τ, Ω(τ)−Ω(0)<<Ω(0).\nΩ(τ) = Ω(0)+dΩ\ndt|0τ\nΩ(τ)−Ω(0) =dΩ\ndt|0τ→dΩ\ndt|0τ <<Ω(0).(24)\nThe adiabatic limit in this case is the limit of fast precession so one can ch oose as the small time τthe (approximate)\nprecession period around the magnetic field, τ=2π\nΩ. The condition becomes\n2π\nΩdΩ\ndt<<Ω. (25)11\nFIG. 3: Spin splitting of energy spectrum in the Rashba model . The green arrows represent the direction of the momentum\nand the red arrows represent the direction of the spin.\nConsider the case in which a small damping term exists α(ds\ndt×s). The equations of motion for the three spin\ncomponents are\ndsx\ndt=−szΩy+α(dsy\ndtsz−dsz\ndtsy)\ndsy\ndt=szΩx+α(dsz\ndtsx−dsx\ndtsz)\ndsz\ndt=sxΩy−syΩx+α(dsx\ndtsy−dsy\ndtsx).(26)\nIt is clear that all the changes in sxare at least of second order in small quantities, so sxcan be treated as a constant.\nThe solution is approximately\nsy(t) =ǫsx\nΩ2x(Ωxt−e−t/τsinΩxt)\nsz(t) =ǫsx\nΩ2[1−e−t/τ(cosΩt−1\nΩτsinΩt)].(27)\nwhereτ=1\nαΩxsx. After the oscillations in sy,szare damped, what remains is\nsy(t >> τ) =ǫsx\nΩxt\nsz(t >> τ) =ǫsx\nΩ2.(28)\nThe spin, which was originally in the plane, acquires a small steady-sta te out-of-plane component. This component\ndepends on sx, andsxhas the opposite sign on different sides of the Fermi surface. As a r esult,szup and down spins\ntravel in opposite directions and a spin-Hall current is established. This derivation, illustrated by Fig. 3, is another\nway to clarify the role of spin precession in steady-state band stru cture spin currents.\nIV. EXTRINSIC SPIN DENSITIES AND CURRENTS\nUntil now the scattering potential has been treated as a scalar. H owever, in general the spin-orbit interaction makes\na contribution to the disorder potential. The disorder potential inc luding the scalar and spin-orbit parts takes the\nform\nUkk′= [1+iλσ·(k×k′)]Ukk′\nV. (29)12\nFIG. 4: An electric field displaces the Fermi surface and the e lectron spins are tilted up for py>0 and down for py<0.\nwithUkk′the matrix element of the scalarpart of the potential. The second term comes from spin-orbit coup ling\ninUkk′andλis a material-specific constant. Configurational averages of term s of the form Uimp\nkk′Uimp\nk′k′′Uimp\nk′′kare also\nlinear in nibut are rather lengthy to be displayed explicitly.\nThe spin-orbit interaction also produces a correction[92] to the po sition operator ˆr=ˆrord+2λσ×kwhereˆrord\nis the ordinary position operator. Since the external electric field e nters through the Hamiltonian HE=eE·ˆrthe\ncorrectionto ˆrgivesaside jump termHj\nE= 2eλσ·k×Eand this side jump term makes an additional contribution[92]\nto the velocity operator vj\nE=−2eλ\n/planckover2pi1σ×E. In the kinetic equation extra spin-dependent driving term −(i//planckover2pi1)[Hj\nE,f0k]\nmust be taken into account due to the change in the position operat or. This term is nonzero if the equilibrium density\nmatrixf0kis spin-dependent. The side jump mechanism is typically classified as ex trinsic although its contributions\nare manifold. These mechanisms were studied extensively in the anom alous Hall effect [89, 90, 91, 92] and recently\ntheir role in steady-state spin densities and currents have been hig hlighted in semiconductors[77, 78, 79, 83] and\nmetals.[84]\nThe solution for nEkis found as above and the spin-dependent scattering term ˆJsacts onnEkand produces an\nadditional source terms for SEk, such that the RHS of Eq. (20) becomes\nΣEk−ˆJs(nEk)−i\n/planckover2pi1[Hj\nE,S0k]−ˆJj(f0k). (30)\nSkew scattering emerges in the second Born approximation as a thir d-order term in the potential U. Substituting for\nnEkand introducing the angles γ1=θ′−θ,γ2=θ′′−θ′,γ3=θ−θ′′and the solid angles ω′,ω′′,\nˆJs(nk) =−6π2λnieτpm∗2k2d−2\n/planckover2pi16δ(k−kF)σ·ˆk×I\nI=/integraldisplaydω′\n(2π)d/integraldisplaydω′′\n(2π)dU(γ1)U(γ2)U(γ3)(ˆk′)E·(ˆk′−ˆk′′).(31)\nThe integraloversolidangles Iis independent of ˆk. In twodimensions the integrandis expanded in Fourierharmonics\nand the integration is straightforward. In three dimensions the int egrand is expanded in Legendre polynomials of γ1,\nγ2andγ3, and the integral over the two solid angles of Pl(γ1)Pm(γ2)Pn(γ3) is independent of ˆk. The modification of\nthe position operator produces an additional side jump scattering term\n−ˆJj(f0k) =−Hj\nEδ(εk−εF)\nτp. (32)\nThis term is related quite literally to transverse jumps undergone by a carrier during scattering [92] and its form\nreflects the conservation of momentum during such a side jump sca ttering event. [79, 92]13\nThe intrinsic source Σ Ekwas discussed in the previous section. The extrinsic source is hence forth denoted by\nTEk=−i\n/planckover2pi1[Hj\nE,S0k]−��Jj(f0k)−ˆJs(nEk). If the band structure spin-orbit interactions are zero one find s easily\nSext\nEk=Hj\nEδ(εk−εF)−ˆJs(nEk)τp, (33)\nwhere the superscript extindicates that only extrinsic contributions are considered. This exp ression averages to zero\nover directions in momentum space and does not give a spin density. I t does however contribute to the spin current,\ngiving a spin-Hall conductivity ∝n−1\nidue to skew scattering and a spin-Hall conductivity neλindependent of nidue\nto side jump. These terms were discussed by Engel et al.[77] and by Hankiewicz and Vignale [79].\nIf band structure spin-orbit interactions are nonzero then, as w as pointed out by several groups,[80, 81, 82, 83]\nspin precession is crucial in establishing the steady state. To illustra te this consider once again the decomposition\nSext\nEk=Sext\nEk/bardbl+Sext\nEk⊥into linearly independent components. In two-dimensional systems grown along (001) the entire\nextrinsic source term is orthogonal to Hso, therefore TEk/bardbl= 0 and there is no term in Sext\nEkthat is∝n−1\ni. The\nsolution is\nSext\nEk⊥=σ·ˆΩk×TEk\n2Ωk. (34)\nThe contributions to Sext\nEk⊥due to skew scattering and side jump contain kto an even power, therefore skew scattering\nand side jump do not contribute to the spin current in the presence of spin precession (they can be restored by an\nexternal magnetic field B∝bardblˆzas found in Ref. [[80]].) The only contribution to the spin current comes from the\nspin-dependent driving term. The spin-Hall conductivity originating from this term is neλregardless of the form of\nthe band structure spin-orbit interaction. Skew scattering in the presence of band structure spin-orbit interactions\ndoes give a steady-state spin density, which was pointed out in [[83]], a nd a similar spin density arises from side jump.\nIt is therefore seen that skew scattering and side jump contribut e very differently to the steady-state density matrix\nwhen band structure spin-orbit interactions are present. This ca n be understood from the following argument. In\norder to get a sizable spin current from skew scattering, spins mus t be conserved as they travel. Scattering processes\nproduce a separation between spin-up and spin-down, while preces sion tries to destroy the orientation of the spins.\nThus once the spins are scattered it is the conserved spin fraction that carries the spin current. Evidently the same\nargument applies for side jumps undergone during scattering.\nV. DEFINITIONS OF THE SPIN CURRENT AND BOUNDARY CONDITIONS\nThe spin current is defined in an intuitive manner by ˆJσ\ni= (1/2){sσ,vi}, and this definition is used by the majority\nof researchers working on this topic. Nevertheless, in the presen ce of band structure spin-orbit interactions this spin\ncurrent is not conserved. The equation of continuity satisfied by t he spin density and current was shown by Shi et al.\nto be different [32] from the usual equation of continuity for the ch arge density and current. A source term exists in\nthis equation which reflects spin non-conservation. As a result, Sh iet al.as well as Bryksin and Kleinert [34] have\nproposed an alternative definition of the spin current accordingto whichˆJσ\ni=d/dt(ˆriˆsσ). The equation of continuity\nsatisfied by this current still contains a source term, but this sour ce term vanishes in many commonly used models.\nThis definition was used in Refs. [[33, 69]]. Recently, Tokatly [35] has a rgued that the intuitive definition of the spin\ncurrent represents a dissipative current which is conjugate to an effective SU(2) electric field.\nIn order to obtain the spin accumulation the kinetic equation needs t o be supplemented by boundary conditions.\nThis was done by a number of groups. [38, 39, 40, 41, 42, 43, 44, 45 ] Unfortunately, it was shown that the form of\nthe spin accumulation depends on the theoretical model of the bou ndary. [43, 44] The nontrivial physics associated\nwith boundary conditions and the implications of various formulations of boundary conditions for spin transport were\ndiscussed by Bleibaum [39] and Tserkovnyak et al.[38] An interesting argument due to Adagideli and Bauer [45] points\nout that, in systems in which the band structure spin current is exp ected to vanish in the bulk (i.e. for Rashba and\nDresselhaus spin-orbit coupling), the situation is not the same near the edges and near metal contacts, where a finite\nspin current exists which can be detected.\nVI. CRYSTAL SYMMETRY\nSpin densities and spin currents in a crystal are closely tied to the sy mmetry of the underlying lattice. An analysis\nrelating the response tensor to the symmetry of the underlying cr ystal lattice has been enlightening in the context of\nnonequilibrium spin densities excited by an electric field. If the respon se of the spin density sto an electric field Eis14\ngiven by sσ=Qσ\njEj, nonzero components for the material-specific spin density respo nse tensor Qσ\njare permitted only\nin gyrotropiccrystals [5]. For spin transport such an analysis was pe rformed in Ref.[[36]], determining the components\nof the spin-current response tensor allowed by symmetry in an elec tric field and providing systematic proof that spin\ncurrents in response to an electric field can be much more complex th an the spin-Hall effect [49]. This result is\ncompletely general and is not sensitive to the definition of the spin cu rrent or to whether the electric field is constant\nor time-dependent.\nThe spin current operator can be defined as ˆJσ\ni=1\n2(ˆsσˆvi+ ˆviˆsσ) orˆJσ\ni=d/dt(ˆriˆsσ). From a symmetry point of\nview these two definitions are equivalent. The spin current ˆJis a second rank tensor that can be decomposed into a\npseudoscalar part, an antisymmetric (spin-Hall) part, and a symme tric part. The pseudoscalar part is tr( ˆJ) =1\n3ˆs·ˆv\nand represents a spin flowing in the direction in which it is oriented. The symmetric and antisymmetric parts are\ngiven, respectively, by1\n2(ˆsσˆvi±ˆvσˆsi). The pseudoscalar and symmetric parts will be referred to as non-spin-Hall\ncurrents. Under the full orthogonalgroup only the antisymmetr ic (spin-Hall) components of ˆJare allowed, indicating\nthat these components are always permitted by symmetry.\nIn general, the spin current responseof a crystal to an electric fi eldEis characterizedby a materialtensor Tdefined\nbyJσ\ni=Tσ\nijEj. Forthe 32crystallographicpoint groupsthe symmetry analysis[10 0] for the tensor Tis establishedby\nmeans of standard compatibility relations [101]. One is particularly inte rested in those groups in which non-spin-Hall\ncomponents may be present. The pseudoscalar part of the spin cu rrent is only allowed by 13 point groups, while the\nsymmetric part is allowed in all systems except those with point group sO,Td(zinc blende), or Oh(diamond). Lower\nsymmetries, allowing non-spin-Hall currents, are characteristic o f systems of reduced dimensionality. In such systems\nit was shown that non-spin Hall currents exist.[36] For a quantum we ll grown along (113), the effective magnetic field\ncorresponding to the Rashba spin-orbit interaction has a compone nt in the ˆz-direction as well, and Ref. [[36]] showed\nthat the spin conductivities σx\nxx,σx\nyy,σz\nyxandσz\nxyare nonzero.\nSkew scattering and side jump contributions to the spin current in t he presence of band structure spin-orbit\ninteractions can also be restored by growing the structure along a direction that is not one of the main crystal axes.\nThe contribution to the spin density matrix due to extrinsic mechanis ms has two parts,\nSext\nEk/bardbl=1\n2/parenleftbiggTEkzΩkz\nΩ2\nk/parenrightbigg\nσ·Ωkτ (35)\nandSext\nEk⊥, which is given by Eq. (34). Taking again a quantum well grown along (1 13),Sext\nEk/bardblgives no spin density\nbut gives a spin current from skew scattering and side jump. Skew s cattering and side jump give the same nonzero\ncomponents of the spin conductivity tensor, σx\nxx,σx\nyy,σz\nyxandσz\nxy.\nVII. EXPERIMENTAL SITUATION\nSpin densities in semiconductors were first predicted by Ivchenko [4 ] and observed the following year in tellurium\nby Vorob’ev. [5] Recent experiments[10, 11, 12] have also reporte d the observation of a steady-state spin density in\nsemiconductors. Experimentally the spin density can be found by Ke rr rotation, in which a beam of linearly-polarized\nlight is sent into the sample and its polarizationvectorrotatesby an a nglethat is proportionalto the spin polarization.\nIn so far as measuring a spin current, the situation is more complicat ed. Initially the approach adopted was to\ngenerate a spin current which would flow to the edges of the sample w here it would produce a spin accumulation.\nThis spin density at the edge of the sample could then be detected by Kerr rotation or an equivalent technique.\nJ. Wunderlich et al.[18] used photoluminescence to detect a spin accumulation due to a s pin-Hall current in a two\ndimensional hole gas, which is believed to be due to spin-orbit interact ions in the band structure. Kato et al.[19]\nused Kerr rotation to detect a spin accumulation due to a spin-Hall c urrent in n-GaAs, in which the spin current is\nbelieved to be due to extrinsic mechanisms [77]. A similar experiment was carried out shortly thereafter by Sih et al.\n[20]. Stern et al.[21] used the same technique to detect a spin-Hall current in ZnSe a t room temperature. Following\nthis work, Sih et al.[22] performed an experiment designed to demonstrate explicitly th at the observed edge spin\naccumulation was due to the spin-Hall effect, as opposed to edge eff ects associated with the electric field. The group\nmanufactured samples with a series of transverse channels, such that the spin-Hall current generated by the electric\nfield was allowed to drift into regions where the electric field was effect ively zero. Since a spin accumulation was still\nmeasured at the edge, this showed unambiguously that the effect w as due to the spin current. Further experiments\nby Stern et al.[23] imaged the spatial distribution of the spin accumulation genera ted by the spin-Hall effect as well\nas its behavior in a magnetic field. Chang et al.[24] also reported, using photoluminescence, a spin accumulation du e\nto the spin-Hall effect in InGaN/GaN superlattices.\nA second type of experiments relies on an effect referred to as the inverse spin-Hall effect. Briefly, a spin current in\nturn generates a charge current, which can be detected by conv entional means. This was shown by Hirsch [14] and its15\nextension to a Landauer-Buttiker type multi-terminal nonlocal me asurement was presented by Hankiewicz et al.[74]\nThis technique was used by Valenzuela and Tinkham [25] to detect a c harge current as a result of the spin current\nin aluminium, and by Saitoh et al.[27] in platinum. Kimura et al.[28] observed the spin-Hall effect in platinum at\nroom temperature and their findings were explained theoretically by Guoet al., [76] who demonstrated that the effect\nis due to band structure spin-orbit interactions. Vila et al.also obtained results for platinum nanowires [29] while\nSekiet al.[30] reported a giant spin Hall effect in FePt/Au devices. The group u sed a multi-terminal device with a\ngold cross and FePt acting as a spin injector. The spin-Hall resistan ce was measured to be 2.9 mΩ and is attributed\nto the large skew scattering in gold, being thus extrinsic in nature. W enget al.[31] carried out similar experiments\non platinum, aluminium and gold and obtained results in agreement with t hose found to date. It should be noted\nthat Cui et al.[26] also observed an electrical current induced by an optically-gen erated spin current.\nVIII. FUTURE DIRECTIONS\nWhereas the scientific community working on steady-state spin den sities and currents appears to be in agreement\nthat spin currents exist and are experimentally measurable, a numb er of questions remain to be addressed in the\nfuture. For example, the relative magnitude of band structure sp in currents and spin currents due to extrinsic\nmechanisms such as skew scattering and side jump remains to be det ermined for a general band structure spin-\norbit interaction. In addition, the fact that no unique definition of t he spin current exists causes difficulties in the\ncomparison of experimental data with theoretical predictions. Th is ambiguity is exacerbated by the fact that different\ndefinitions of the spin current give results that often differ by a sign . [32, 33] Thanks to the non-conservation of\nspin, the relationship between spin current and spin accumulation at the boundary remains to be clarified. It appears\nthat what happens at the boundary is sensitive to the type of boun dary conditions assumed.[43, 44] Thus so far as\nquantitative interpretation of experimental data is concerned, t heory still has some way to go.\nOn the experimental side, despite tremendous progress, the com munity is still searching for a reliable way to\nmeasure, as opposed to detect, spin currents directly. A possible new experimental avenue relies o n magnetoresistance\ncaused by an edge spin accumulation, proposed by Dyakonov. [85] A spin current produces a spin accumulation near\nthe sample edges, which in turn causes the sample resistance to dec rease by a small amount, whereas an external\nmagnetic field can destroy the edge spin polarization and yield a positiv e magnetoresistance. An alternative path was\nfollowed by Wang et al.,[46] who started from the Dirac equation and obtained in the weakly relativistic limit a set of\nMaxwell equations in the presence of spin-orbit interactions. Altho ugh the spin current does not appear explicitly, it\nis contained in the Maxwell equations and the authors demonstrate that the relativistic conservation laws imply that\nthe spin current yields an electrical polarization, which could be dete cted directly. Similarly, a set of Maxwell-like\nequations for spin was formulated by Bernevig et al.[47] These methods remains to be attempted experimentally.\nPractically, the question of what to do with the electron spin once it h as been generated or transported to the\nedge of the sample remains. The revolutionary electronic device tha t harnesses spin currents for a practical purpose\nremains to be made, and the challenge of its design confronts exper imentalists and theorists alike.\nThe research at Argonne National Laboratory was supported by the US Department of Energy, Office of Science,\nOffice of Basic Energy Sciences, under Contract No. DE-AC02-06C H11357.\n[1] F. Meier and B. P. Zakharchenya, Optical Orientation (Elsevier, Amsterdam, 1984.)\n[2] G. Schmidt et al., Phys. Rev. B 62, R4790(2000).\n[3] T. Jungwirth, J. Sinova, J. Masek, J. Kucera, and A. H. Mac Donald, Rev. Mod. Phys. 78, 809 (2006).\n[4] E. L. Ivchenko and G. E. Pikus, JETP Lett. 27, 604 (1978).\n[5] L. E. Vorob’ev, E. L. Ivchenko, G. E. Pikus, I. I. Farbstei n, V. A. Shalygin, and A. V. Sturbin, JETP Lett. 29, 441\n(1979).\n[6] L. S. Levitov, Y. V. Nazarov, and G. M. Eliashberg, Sov. Ph ys. JETP 61(1), 133 (1985).\n[7] V. M. Edelstein, Solid State Comm. 73, 233 (1990).\n[8] A. G. Aronov, Y. B. Lyanda-Geller, and G. E. Pikus, Sov. Ph ys. JETP 73, 537 (1991).\n[9] L. I. Magarill, A. V. Chaplik, and M. V. Entin, Semiconduc tors35(9), 1081 (2001).\n[10] Y. K. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom , Phys. Rev. Lett. 93, 176601 (2004).\n[11] S. Ganichev et al, J. Magn. Magn. Mater. 300, 127 (2006).\n[12] A. Y. Silov, P. A. Blajnov, J. H. Wolter, R. Hey, K. H. Ploo g, and N. S. Averkiev, Appl. Phys. Lett. 85, 5929 (2004).\n[13] M. I. D’yakonov and V. I. Perel’, Zh. Eksp. Teor. Fiz 60, 1954 (1971).\n[14] J. E. Hirsch, Phys. Rev. Lett. 83, 1834 (1999).\n[15] S. Murakami, N. Nagaosa, and S.-C. Zhang, Science 301, 1348 (2003).\n[16] J. Sinova, D. Culcer, Q. Niu, N. A. Sinitsyn, T. Jungwirt h, and A. H. MacDonald, Phys. Rev. Lett. 92, 126603 (2004).16\n[17] S. A. Crooker and D. L. Smith, Phys. Rev. Lett. 94, 236601 (2005).\n[18] J. Wunderlich, B. Kaestner, J. Sinova, and T. Jungwirth , Phys. Rev. Lett. 94, 047204 (2005).\n[19] Y. K. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom , Science 306, 1910 (2004).\n[20] V. Sih, R. C. Myers, Y. K. Kato, W. H. Lau, A. C. Gossard, an d D. D. Awschalom, Nature Phys. 1, 31-35 (2005).\n[21] N. P. Stern, S. Ghosh, G. Xiang, M. Zhu, N. Samarth, and D. D. Awschalom, Phys. Rev. Lett. 97, 126603 (2006).\n[22] V. Sih, W.H. Lau, R.C. Myers, V.R. Horowitz, A.C. Gossar d, D.D. Awschalom, Phys. Rev. Lett. 97, 096605 (2006).\n[23] N. P. Stern, D. W. Steuerman, S. Mack, A. C. Gossard, D. D. Awschalom, Appl. Phys. Lett. 91, 062109 (2007).\n[24] H. J. Chang et al, Phys. Rev. Lett. 98, 136403 (2007).\n[25] S. O. Valenzuela and M. Tinkham, Nature 442, 176 (2006).\n[26] X.-D. Cui, S.-Q. Shen, J. Li, W. Ge, and F.-C. Zhang, Appl . Phys. Lett. 90, 242115 (2007).\n[27] E. Saitoh et al, Appl. Phys. Lett. 88, 182509 (2006).\n[28] T. Kimura et al, Phys. Rev. Lett. 98, 156601 (2007).\n[29] L. Vila et al, Phys. Rev. Lett. 99, 226604 (2007).\n[30] T. Seki, Y. Hasegawa, S. Mitani, S. Takahashi, H. Imamur a, S. Maekawa, J. Nitta, and K. Takanashi, Nature Materials\n7, 125 - 129 (2008).\n[31] K. C. Weng, N. Chandrasekhar, C. Miniatura, B.-G. Engle rt, arXiv:0804.0096.\n[32] J. Shi, P. Zhang, D. Xiao, and Q. Niu, Phys. Rev. Lett. 96, 076604 (2006).\n[33] N. Sugimoto, S. Onoda, S. Murakami, and N. Nagaosa, Phys . Rev. B 73, 113305 (2006).\n[34] V. V. Bryksin and P. Kleinert, Phys. Rev. B 73, 165313 (2006).\n[35] I. V. Tokatly, Phys. Rev. Lett. 101, 106601 (2008).\n[36] D. Culcer and R. Winkler, Phys. Rev. Lett. 99, 226601 (2007).\n[37] E. I. Rashba, Phys. Rev. B 70, 201309 (2004).\n[38] Y. Tserkovnyak, B. I. Halperin, A. A. Kovalev, and A. Bra taas, Phys. Rev. B 76, 085319 (2007).\n[39] O. Bleibaum, Phys. Rev. B 74, 113309 (2006).\n[40] G. Usaj and C. A. Balseiro, Europhys. Lett. 72, 631 (2005).\n[41] A. G. Mal’shukov, L. Y. Wang, C. S. Chu, and K. A. Chao, Phy s. Rev. Lett. 95, 146601 (2005).\n[42] K. Nomura, J. Wunderlich, J. Sinova, B. Kaestner, A. H. M acDonald, and T. Jungwirth, Phys. Rev. B 72, 245330 (2005).\n[43] Tse, W.-K., Fabian, J., Zutic, I. and Das Sarma, S. (2005 ) Spin accumulation in the extrinsic spin-Hall effect, Phys. Rev.\nB72, 241303.\n[44] V. M. Galitski, A. A. Burkov, and S. Das Sarma, Phys. Rev. B74, 115331 (2006).\n[45] I. Adagideli and G. E. W. Bauer, Phys. Rev. Lett. 95, 256602 (2005).\n[46] Y. Wang, K. Xia, Z.-B. Su, and Z. Ma, Phys. Rev. Lett. 96, 066601 (2006).\n[47] B. A. Bernevig, X. Yu, and S.-C. Zhang, Phys. Rev. Lett. 95, 076602 (2005).\n[48] D. Culcer and R. Winkler, Phys. Rev. B 76, 245322 (2007).\n[49] N. A. Sinitsyn, E. M. Hankiewicz, W. Teizer, and J. Sinov a, Phys. Rev. B 70, 081312 (2004).\n[50] J. Inoue, G. E. W. Bauer, and L. W. Molenkamp, Phys. Rev. B 70, 041303 (2004).\n[51] O. V. Dimitrova, Phys. Rev. B 71, 245327 (2005).\n[52] P. Schwab and R. Raimondi, Eur. Phys. J. B 25, 483 (2002).\n[53] E. G. Mishchenko, A. V. Shytov, and B. I. Halperin, Phys. Rev. Lett. 93, 226602 (2004).\n[54] C. Grimaldi, E. Cappelluti, and F. Marsiglio, Phys. Rev . B.73, 081303 (2006).\n[55] S. Y. Liu, X. L. Lei, and N. J. M. Horing, Phys. Rev. B 73, 035323 (2006).\n[56] A. Khaetskii, Phys. Rev. Lett. 96, 056602 (2006).\n[57] A. V. Shytov, E. G. Mishchenko, H.-A. Engel, and B. I. Hal perin, Phys. Rev. B 73, 075316 (2006).\n[58] A. G. Mal’shukov and K. A. Chao, Phys. Rev. B 71, 121308 (2005).\n[59] S. Zhang and Z. Yang, Phys. Rev. Lett. 94, 066602 (2005).\n[60] L. Sheng, D. N. Sheng, C. S. Ting, and F. D. M. Haldane, Phy s. Rev. Lett. 95, 136602 (2005).\n[61] T. Damker, H. Bottger, and V. Bryksin, Phys. Rev. B 69, 205327 (2004).\n[62] B. K. Nikolic, L. P. Zarbo, and S. Souma, Phys. Rev. B 73, 075303 (2006).\n[63] O. Bleibaum and S. Wachsmuth, Phys. Rev. B 74, 195330 (2006).\n[64] J. Schliemann and D. Loss, Phys. Rev. B 71, 085308 (2005).\n[65] P. -Q. Jin and Y. -Q. Li, Phys. Rev. B 74, 085315 (2006).\n[66] J. Li, L. Hu, and S. Q. Shen, Phys. Rev. B 71, 241305 (2005).\n[67] S. Y. Liu and X. L. Lei, Phys. Rev. B 72, 155314 (2005).\n[68] P. L. Krotkov and S. Das Sarma, Phys. Rev. B 73, 195307 (2006).\n[69] M. Zarea. and S. Ulloa, Phys. Rev. B. 73, 165306 (2006).\n[70] N. A. Sinitsyn, A. H. MacDonald, T. Jungwirth, V. K. Duga ev, and J. Sinova, Phys. Rev. B 75, 045315 (2007).\n[71] D. N. Sheng, L. Sheng, Z. Y. Weng, and F. D. M. Haldane, Phy s. Rev. B 72, 153307 (2005).\n[72] G. Y. Guo, Y. Yao, and Q. Niu, Phys. Rev. Lett. 94, 226601 (2005).\n[73] Y. Yao and Z. Fang, Phys. Rev. Lett. 95, 156601 (2005).\n[74] E. M. Hankiewicz, L. W. Molenkamp, T. Jungwirth, and J. S inova, Phys. Rev. B 70, 241301 (2004).\n[75] R. Raimondi et al., Phys. Rev. B 74, 035340 (2006).\n[76] G. Y. Guo, S. Murakami, T.-W. Chen, and N. Nagaosa, Phys. Rev. Lett. 100, 096401 (2008).\n[77] H. -A. Engel, B. I. Halperin, and E. I. Rashba, Phys. Rev. Lett.95, 166605 (2005).\n[78] S. Onoda, N. Sugimoto, and N. Nagaosa, Prog. Theor. Phys .116, 61 (2006).17\n[79] E. M. Hankiewicz and G. Vignale, Phys. Rev. B 73, 115339 (2006).\n[80] E. M. Hankiewicz and G. Vignale, Phys. Rev. Lett. 100, 026602 (2008).\n[81] W.-K. Tse and S. Das Sarma, Phys. Rev. B 74, 245309 (2006).\n[82] L. Hu, Z. Huang, and S. Hu, Phys. Rev. B 73, 235314 (2006).\n[83] J. L. Cheng and M. W. Wu, J. Phys. Cond. Mat. 20, 085209 (2008).\n[84] R. V. Shchelushkin and A. Brataas, Phys. Rev. B 71, 045123 (2005).\n[85] M. I. Dyakonov, Phys. Rev. Lett. 99, 126601 (2007).\n[86] G. Dresselhaus, Phys. Rev. 100, 580 (1955).\n[87] Y. A. Bychkov and E. I. Rashba, JETP Lett. 39, 66 (1984).\n[88] J. M. Luttinger, Phys. Rev. 102, 1030 (1956).\n[89] Q. Kohn and J. M. Luttinger Phys. Rev. 108, 590 (1957).\n[90] J. Smit, Physica 21, 877 (1955); ibid. 23, 39 (1958).\n[91] L. Berger, Phys. Rev. B 2, 4559 (1970).\n[92] P. Nozieres and C. Lewiner, Le Journal de Physique 34, 901 (1973).\n[93] C. L. Kane and E J. Mele, Phys. Rev. Lett. 95, 146802 (2005).\n[94] S. Murakami, N. Nagaosa, and S.-C. Zhang, Phys. Rev. Let t.93, 156804 (2004).\n[95] R. Winkler, Phys. Rev. B 69, 045317 (2004).\n[96] Out of the many texts on this topic, the excellent exposi tion by J. Rammer, Quantum Transport Theory (Perseus Books,\nReading MA, 1998) stands out. Chs. 2, 3, 7 and 8 discuss the fun damentals needed to understand linear response theory\nin systems with disorder.\n[97] G. Sundaram and Q. Niu, Phys. Rev. B 5914915 (1999).\n[98] D. Culcer, Y. Yao, A. H. MacDonald, and Q. Niu, Phys. Rev. B 72, 045215 (2005).\n[99] J. J. Sakurai, Modern Quantum Mechanics (Addison-Wesley, Reading MA, 1994).\n[100] G. L. Bir and G. E. Pikus, Symmetry and Strain-Induced Effects in Semiconductors (Wiley, New York, 1974).\n[101] G. F. Koster, J. O. Dimmock, R. G Wheeler, and H. Statz, Properties of the Thirty-two Point Groups (MIT Press,\nCambridge MA, 1963)." }, { "title": "1305.2443v3.Spin_dynamics_in_a_spin_orbit_coupled_Fermi_gas.pdf", "content": "arXiv:1305.2443v3 [cond-mat.quant-gas] 26 Sep 2013Spin dynamics in a spin-orbit coupled Fermi gas\nStefan S. Natu∗and S. Das Sarma\nCondensed Matter Theory Center and Joint Quantum Institute , Department of Physics,\nUniversity of Maryland, College Park, Maryland 20742-4111 USA\nWe study the dynamics of a non-degenerate, harmonically tra pped Fermi gas following a sudden\nramp of the spin-orbit coupling strength using a Boltzmann e quation approach. In the absence of\ninteractions and a Zeeman field, we solve the spin-orbit coup led Boltzmann equation analytically ,\nand derive expressions for the phase-space and temporal dyn amics of an arbitrary initial spin state.\nFor a fully spin polarized initial state, the total magnetiz ation exhibits collapse and revival dynamics\nin time with a period set by the trapping potential. In real sp ace, this corresponds to oscillations\nbetween a fully polarized state and a spin helix. To make pred ictions relevant to current experiments\non spin-orbit coupled Fermi gases, we then numerically stud y the dynamics in the presence of an\nadditional momentum independent Zeeman field. We find that th e spin helix is robust for weak\nmagnetic fields but disappears for stronger field strengths. Finally, we explore the spin dynamics\nin the presence of interactions and find that weak interactio nsenhance the amplitude of the spin\nhelix.\nI. INTRODUCTION\nThephysicsofspin-orbitcouplingisattheheartoffun-\ndamental phenomena such as the spin Hall effect [1, 2]\nas well as practical devices such as the spin transistor\n[3]. Key to these developments in the field of spintron-\nics is the understanding of how parameters such as the\nspin-orbit coupling, interaction, disorder, and geometry\nseparately, as well as collectively influence spin dynamics\n[4,5]. Thecreationoflowtemperatureatomicandmolec-\nular spin-orbit coupled Bose and Fermi gases has paved\nthe way for studying this physics in a setting where these\nparameters are well characterized [6–10]. Furthermore, a\ntool largely unique to ultra-cold gases is the ability to\ninduce the spin-orbit coupling dynamically , thereby en-\nabling the study of out of equilibrium physics in these\nsystems. In this paper, we solve the dynamics of a non-\ninteracting non-degenerate Fermi gas following a sudden\nramp of the spin-orbit coupling strength. The resulting\nout-of-equilibriumdynamicsisrich[11–13]: coherencein-\nherent in cold atomic gases, but almost always absent in\nsolid state systems leads to collapse and revival of the\ntotal magnetization [12]. In real space, this is manifested\nby the spontaneous appearance of a helical spin texture\nwhichistheanalogueofthepersistentspinhelixobserved\nin two-dimensional electron gases [14, 15].\nA further advantage of cold atomic systems is the abil-\nity to use magnetic fields to control the interactions be-\ntween the atomic states [16, 17]. This, particularly in\nthe presenceofspin-orbitcouplingands-wavesuperfluid-\nity, may enable experimentalists to realize novel states of\nmatter with topological properties whose excitations ex-\nhibit non-Abelian statistics [18, 19]. In addition to their\nfundamental importance, such topological states of mat-\nter may serve as a platform for fault tolerant quantum\ncomputation [20]. However a key challenge in observing\n∗Electronic address: snatu@umd.eduthis physicsin ultra-coldgasesisthe inherent difficulty in\nattaining sufficiently low temperatures needed to realize\nthese topological states. In contrast, the non-equilibrium\ndynamics we study here occurs at high temperatures and\ncan be observed in currentexperiments. We numerically\nstudy the spin dynamics of a fully polarized, weaklyin-\nteracting Fermi gas using a collisionlessBoltzmann equa-\ntion. We find large amplitude spin waves analogous to\nthose previously observed in dilute spin polarized Hy-\ndrogen [21–25] and more recently in ultra-cold Bose and\nFermi gases [26–31]. Our numerical study of spin dy-\nnamics in the collisionless spin-orbit coupled Fermi gas\ncomplements a recent, linear response study on a homo-\ngeneous system by Tokatly and Sherman [11]. We point\nout that the ultra-low temperature degenerate version of\nour system (with spin-orbit coupling and large Zeeman\nsplitting) would manifest topological superfluidity in the\npresenceofordinarys-wavesuperfluidityinducedbysuit-\nable Feshbach resonance [19].\nWeshowthatspin-orbitcoupling, whencombinedwith\nthe long coherence times inherent to cold atomic and\nmolecular gases, leads to surprising dynamical phenom-\nena even at high temperatures where the cold gas is not\nnecessarily quantum degenerate. Furthermore the non-\ndegenerate gas is an ideal conceptual starting point for\nstudying how interactions influence the spin dynamics in\nthe presence of spin-orbit coupling.\nThe effects predicted in our work are unlikely to be of\nmuch experimental significance in solid state spin-orbit\ncoupled systems because of strong disorder and decoher-\nence intrinsically present in these systems as well as the\nultra-fast time scales for spin relaxation [32]. But in\ncold atomic and molecular systems, our proposed physics\ncould be studied in existinglaboratory systems. In prin-\nciple, the physics we predict should be present in both\natomic/molecular Fermi/Bose gases since it is an intrin-\nsically high temperature phenomenon, but in light of re-\ncent experiments [8–10], we will discuss our theoretical\ndetails using the atomic Fermi gas as the representative\nsystem of study.2\nThis paper is organized as follows: In Section II, we\ndescribe our system and derive the collisionless Boltz-\nmann equation for a two-component Fermi gas in a 2D\nharmonic trap in the presence of spin-orbit coupling. In\nthe subsequent sections, we choose an initial state which\nis a stationary state of the Hamiltonian in the absence\nof spin-orbit coupling. We then drive the system out-\nof-equilibrum by suddenly turning on the spin-orbit cou-\npling and study the resulting dynamics. In Section III,\nwe solve the Boltzmann equation in the absence of in-\nteractions, and in Section IV, we consider the effect of\ninteractions on the spin dynamics. We summarize our\nresults in Section V.\nII. SETUP\nIn the absence of spin-orbit coupling, the Hamiltonian\nfortwohyperfinestatesofaFermigascanbeexpressedas\na sum of single particle and two-body interaction terms:\nH=Hs+Hint (1)\nwhere the single particle Hamiltonian is composed of a\nkinetic term and a potential term arising from the exter-\nnal trapping potential\nHs=/summationdisplay\ni=↑,↓/integraldisplay\nd3rψ†\ni/parenleftBig\n−/planckover2pi12∇2\nr\n2m+U(r)/parenrightBig\nψi\nwhereψidenotes the fermionic annihilation operator for\na particle in hyperfine state {↑,↓}and massm. Here\nU(r) refers to the external trapping potential, which we\nassume to be cylindrically symmetric U(r) =1\n2mω2\nr(x2+\ny2)+1\n2mω2\nzz2, whereωrandωzare the trapping frequen-\ncies in the radial and longitudinal directions respectively.\nAs the relevant spin-orbit physics is two-dimensional, we\nassume a quasi-two dimensional, pancake geometry, ob-\ntained by tight confinement in the longitudinal ( z) di-\nrection (ωz≫ωr). For simplicity, we assume that both\nthe atomic states experience identical trapping poten-\ntials, but all our calculations can be readily extended to\ninclude more general trapping potentials realized in ex-\nperiments.\nAt the ultra-cold temperatures realized in these exper-\niments, the dominant contribution to scattering comes\nfrom the s-wave channel. The interaction Hamiltonian\ntherefore takes the simple form of a contact interaction:\nHint=g\n2/integraldisplay\ndrψ†\n↑(r)ψ†\n↓(r)ψ↓(r)ψ↑(r) (2)\nwith interaction strength g= 4π/planckover2pi12a/mwhereadenotes\nthe s-wave scattering length. A key advantage of the\nfermonicexperiments[8–10] istheabilityto usemagnetic\nfields to tune the interactions between different internal\nstates, viaaFeshbachresonance[16,17]. Hereweworkin\nthe weakly interacting regime, which can be realized by\nworkingnearthe zerocrossingofthe Feshbachresonance.The spin-orbit Hamiltonian containing terms linear in\nmomentum takes the form:\nHSOC=α(σxpy−σypx)+β(σxpx−σypy) (3)\nwhere the first term is the Rashba contribution\nand the second term is the Dresselhaus contribution,\nparametrized by the coupling constants αandβrespec-\ntively. Here σx,σyandσzdenote Pauli matrices.\nIn the ultra-coldgassetting, spin-orbit couplingis gen-\nerated by using a pair of Raman beams to drive tran-\nsitions between two hyperfine states of an atom, while\nsimultaneously imparting a momentum kick [6–9]. The\nresulting spin-orbit Hamiltonian takes the Rashba equal\nDresshelhaus form α=β, where the magnitude of αis\ndetermined by the wave-length of the Raman beams and\nthe angle at which they intersect. The Raman beams\nalso produce a momentum independent Zeeman field,\nHZ=−/planckover2pi1ΩRσz, where Ω Ris the strength of the Ra-\nman coupling, and is proportional to the intensity of the\nRaman lasers. As we show below, this term plays an\nimportant role in the dynamics. The scheme described\nhere was first successfully demonstrated in experiments\nat NIST using bosonic87Rb [6, 7]. Recently, a simi-\nlar scheme was used to generate spin-orbit coupling in a\nFermi gas of6Li and40K [8–10].\nAs we are primarily motivated by current experiments,\nwe limit ourselves to the case α=βin Eq. 3. In this\nlimit, the spin-orbit Hamiltonian can be diagonalized by\nindependently rotating the momentum co-ordinate and\nperforming a global spin rotation. We will denote the di-\nagonal basis as{ψ+,ψ−}. Throughout, we will use both\nthe diagonal basis ( {+,−}) and the pseudo-spin basis\n({↑,↓}) corresponding to the original hyperfine states,\ndepending on the context. The generalization to α/ne}ationslash=β\nis straightforward within our formalism, and will be the\nsubject of a future work [33].\nAll of the work described here is in the non-degenerate\nlimit. The onlyrequirementonthe temperature Tis that\nit should be smaller than the detuning energy between\nthe magnetic sublevels, so that the gas can be described\nby a two-level (pseudo-spin1\n2) system. This is readily\naccomplished as the splitting between hyperfine levels is\nmuch larger than the Fermi energy for typical densities\n[8].\nThe physics of the weakly interacting gas is domi-\nnated by coherent mean-field dynamics which occurs on\na timescale τmf∼(an0/m)−1which is much faster than\nthe timescale for energy exchanging collisions τcoll∼\n(4πa2n0v)−1(vis the characteristic velocity of the parti-\ncles andn0is the density). For a trapped gas of N∼105\n6Li atoms at a temperature T∼10−6K, the collision-\nless limit (τmf≪τcoll) corresponds to scattering lengths\na∼10aB, whereaBis the Bohr radius. The recent ex-\nperiment of Cheuk et al.is already in this regime [9],\nwhile the experiment of Wang et al.[8] has stronger in-\nteractions of a∼200aB, which can be tuned near zero\nusing a Feshbach resonance [17].3\nMathematically, the weakly interacting gas can be de-\nscribed using a collisionlessBoltzmann equation. Follow-\ningRef. [34], weusetheHeisenbergequationsfor ψσ(r, t)\nto derive the equations of motion for the spin dependent\nWigner function\n← →F=/parenleftbigg\nf↑↑(p,R,t)f↑↓(p,R,t)\nf↓↑(p,R,t)f↓↓(p,R,t)/parenrightbigg\n(4)\nfσσ′(p,R,t) =/integraldisplay\ndreip·r/an}b∇acketle{tψ†\nσ(R−r\n2,t)ψσ′(R+r\n2,t)/an}b∇acket∇i}ht,\nwhich is the quantum analogue of the classical distri-\nbution function. Here prepresents the momentum,\nr=r1−r2is the relative coordinate and R=r1+r2\n2\nis the center of mass coordinate.\nThe diagonal components of← →Fcan be integrated\nin momentum to give the respective spin densities\nnσσ(R,t) =/an}b∇acketle{tψ†\nσ(R,t)ψσ(R,t)/an}b∇acket∇i}ht=/integraltextdp\n(2π)3fσσ(p,R,t),\nwhile the off-diagonal components correspond to quan-\ntum coherences that are absent in a classical model of a\nspin-1\n2gas.\nFor the pancake geometry considered here, all the\nrelevant dynamics is two dimensional. Assuming ther-\nmal equilibrium in the longitudinal direction, we de-\ncompose the Wigner function in the radial and axial\ndirections as fσσ′(p,R,t) =fσσ′(pr,r,t)f(pz,z) where\nf(pz,z) =e−β(p2\nz/2m+1\n2mω2\nzz2)whereβ= 1/kBT.\nWe obtain a two dimensional density n2D\nσσ′(r,t) =/integraltext\ndprfσσ′(pr,r,t)/integraltext\ndpzf(pz,z) by integrating out\nthe longitudinal co-ordinate. Averaging over the\nz−direction, we obtain an effective quasi 2D Boltzmann\nequation:\n∂t← →F+p\nm·∇r← →F−∇rU∇p← →F=i[← →V,← →F]+ (5)\n1\n2{∇r← →V,∇p← →F}+iαp+[σ+,← →F]+α\n2{σ+,∇r+← →F}\nwherep+=px+py,σ+=σx+σy, and the interaction\npotential← →Vis [28]:\n← →V=/parenleftbigg\ng2Dn↓↓−/planckover2pi1ΩR−gn↑↓\n−gn↓↑gn↑↑+/planckover2pi1ΩR/parenrightbigg\n(6)\nwhereg2Dis an effective two-dimensional interaction\nstrength given by g2D= 2√π/planckover2pi12a/(mΛth), where Λ th=/radicalbig\n2π/planckover2pi12/mkBT. The diagonal components in the interac-\ntion matrix arise from forward scattering (Hartree) while\nthe off-diagonal terms arise from exchange interactions\n(Fock). Commutatorsandanti-commutatorsaredenoted\nby [,] and{,}respectively.\nThe single particle limit of Eq. 5 was derived by\nMishchenko and Halperin [5], who used this approach to\nstudy the transport properties of a 2D electron gas. Here\nwe generalize the Boltzmann equation to the include ef-\nfect of← →Von the phase space and spin space evolution of\nthe Wigner function.\nIn general, Eq. 5 is a non-linear matrix equation which\nhas to be solved numerically. Below we assume an initialstatewhichisastationarystateoftheHamiltonianinthe\nabsence of spin-orbit coupling ( α= ΩR= 0). We then\nsuddenly turn on the Raman coupling, and investigate\nthe resulting out-of-equilibrium dynamics.\nIII. NON-INTERACTING LIMIT\nA.ΩR= 0\nThe non-interacting limit in ultra-cold Fermi gases is\nachieved by working at the zero crossing of a Feshbach\nresonance. As the magnetic fields corresponding to the\nzerocrossingofthe interactionsaretypicallysmall[9], we\ndo not expect the Raman couplings to deviate apprecia-\nbly from their zero field values [35]. We first consider the\ncase where upon switching on the Raman coupling, the\nspin-orbit coupling ( α) is non-zero, but the Zeeman term\nΩR= 0. While this scenariodoes not correctlymodel the\npresent experiments, in this limit the Boltzmann equa-\ntion is exactly soluble for an arbitrary initial spin state,\nthus serving as a conceptual starting point. We remark\nthat although we only consider the non-degenerate limit\nhere, our results can be readily generalized to tempera-\ntures below the Fermi temperature.\nIn order to proceed, we introduce dimensionless po-\nsition and momentum coordinates ˜ r=r/rtrapand\n˜p=p√\n2π/planckover2pi1/Λth, wherertrap=/radicalbig\n/planckover2pi1/mωris the char-\nacteristic length scale of motion in the trap and and\nΛth=/radicalbig\n2π/planckover2pi12/mkBTis the thermal deBroglie wave-\nlength. We normalize time in units of the radial trap-\nping frequency ˜t=t/ωr. We also introduce a parameter\nη=/radicalbig\n/planckover2pi1ωr/kBT. As the spin-orbit Hamiltonian only\ncouples to momentum in the px+pydirection, it suf-\nfices to consider the evolution of the distribution in the\np+=px+pyandr+=x+ydirections of phase space.\nConsider an arbitrary initial spin state given by the\nWigner distribution function:← →F(˜p+,˜r+,t= 0) =\ne−1\n4(˜p2\n++η2r2\n+)← →fwhere← →fis a 2×2matrix corresponding\nto the initial spin state. We omit the p−,r−directions\nfor now as they have no dynamics. In the absence of\nspin-orbit coupling or interactions ( α= ΩR=g2D= 0),\nthe initial state is stationary.\nNext, we expressthe Boltzmann equation in the diago-\nnal basis by performing the global transformation← →F→\nB†← →f Bwhere the unitary matrix B=/parenleftBigg1√\n2−1+i\n2\n1+i\n21√\n2/parenrightBigg\nrotatesσx+σyto√\n2σz. In the rotated basis, the spin +\nand−components evolve independently, and the prob-\nlem reduces to a single particle problem in the presence\nof amomentum dependent magnetic field.\nThe dynamics in phase space can now be solved by\nmaking the following ansatz for the diagonal components\nof the rotated spin density matrix:\nF++/−−(˜p+,˜r+,˜t) =A+/−e−1\n4{(˜p+∓a(˜t))2+η2(˜r+∓b(˜t))2}(7)4\n/Minus10/Minus505100Π\n2Π3Π\n22Π\nr/Slash1rtrapt/LParen11/Slash1Ωtrap/RParen1\nFIG. 1: Time evolution of the longitudinal (density plot) an d\ntransverse magnetization (arrows) in the r+=x+ydirection,\nfollowing a sudden ramp of the spin-orbit coupling strength .\nThe Zeeman term is set to zero here (Ω R= 0). Brighter\ncolors indicate positive magnetization ( ↑) and darker colors\nindicate negative magnetization ( ↓). The arrows indicate the\ndirection of the magnetization in the x−yplane, and the\nlength of the arrows indicate the magnitude of the transvers e\nspin normalized to the total spin. At t= 0 all the spins are\npointing in the zdirection (all atoms in the ↑state). spin-\norbit coupling causes the atoms to precess in time, and at\nhalf the trapping period a spin helix is produced. The wave-\nvector of the helix at t=π/ωrdepends only on the spin-orbit\ninteraction and is λsh=π rtrap/2˜α, wherertrap=/radicalbig\n/planckover2pi1/mωr.\nTo clearly illustrate this effect, we choose a weak spin-orbi t\ncoupling of ˜ α=√\n2α/radicalbig\nm//planckover2pi1ωr= 0.125 and η=/radicalbig\n/planckover2pi1ω/kBT=\n0.25 in these simulations. The initial state is recovered afte r\nt= 2π/ωr.\nwherea(0) =b(0) = 0. The coefficients A+/−are the\ndiagonal matrix elements of the spin density matrix← →f\nafter rotation into the {+,−}basis.\nSubstituting the ansatz of Eq. 7 into Eq. 5 we find\na(˜t) = ˜αη(cos(˜t)−1) andb(˜t) = ˜αsin(˜t), where we have\nintroduced a dimensionless, spin-orbit coupling constant\n˜α=√\n2α/radicalbig\nm//planckover2pi1ωr, which parametrizes the strength of\nthe spin-orbit coupling relative to the trapping potential.\nThus the rotated spin densities simply perform oscilla-\ntions in real and momentum space with an amplitude set\nby the strength of the spin-orbit interaction and period\nset by the trap frequency.\nSimilarly, one can solve for the dynamics of the off-\ndiagonal components to find:\nF+−(˜p,+˜r+,˜t) =A+−e−4˜α2(1−cos(˜t))\nη2×(8)\ne−1\n2{(˜p+−2i˜α/ηsin(˜t))2+η2(˜r+−2i˜α/η2(1−cos(˜t)))2}\nwhereA+−is the off-diagonal matrix element of the spin\ndensity matrix after rotation into the ±basis. The dy-\nnamics ofF−+is obtained by replacing ˜ α→−˜αin Eq. 8.\nUnlike the diagonal components, the magnitudes of theoff-diagonal components are notconserved and oscillate\nin time.\nThe corresponding spin densities are found by inte-\ngrating the above expressions for the Wigner functions\nin momentum space. By rotating the diagonalbasis back\ninto the hyperfine basis {↑,↓}, one obtains the dynamics\nof an arbitrary initial spin state.\nTo illustrate the role of quantum coherence, we con-\nsider the dynamics of a fully polarized initial state cor-\nresponding to all particles in the ↑state. We study the\ndynamics of the longitudinal magnetization density and\nthe total magnetization:\nmz(r,t) =/integraldisplay\ndp/parenleftBig\nf↑↑(p,r,t)−f↓↓(p,r,t)/parenrightBig\n(9)\nM(t) =/integraldisplay\ndrmz(r,t)\nThe zero temperature dynamics of the total magneti-\nzation for this initial state was considered previously by\nStanescu, Zhang and Galitski [12]. By exactly solving\nfor the quantum dynamics in a trap, they demonstrated\nthat the total magnetization exhibits collapse and revival\ndynamics, and produced analytic formulas for the total\nmagnetization in weak spin-orbit coupling limit.\nHere we show that similar dynamics also occurs in the\nnon-degenerate gas, which is much more readily acces-\nsible in experiments. Furthermore, we obtain analytic\nexpressions for the total magnetization for arbitrary val-\nues of the spin-orbit coupling.\nRotating the spin polarized state to the diagonal ba-\nsis, one finds that the density matrix has both diagonal\nand off-diagonal matrix elements ( A+=A−= 1/2 and\nA+−=−1−i\n2√\n2), whose dynamics is given by Eqns. (7, 8).\nTransforming back to the hyperfine basis and integrating\nover momentum, the longitudinal magnetization density\ntakes the form:\nmz(˜r+,˜t)∼e−1\n2η2˜r2\n+−˜α2\nη2(1−cos(2˜t))× (10)\ncos(2˜r+˜α(1−cos(˜t))\nand the total magnetization is:\nM(˜t)∼/radicalbigg2π\nηe−4˜α2\nη2(1−cos(˜t))(11)\nwhere we haveignored an overallnormalizationfactor re-\nsultingfromintegrationovermomentum. The expression\nfor the transverse magnetization density is rather cum-\nbersome, but the total transverse magnetization remains\nzero at all times.\nFromEq.11, it isclearthatthe totallongitudinalmag-\nnetization exhibits collapse and revival dynamics in time\nwith a period which depends onlyon the trapping poten-\ntial, and is completely independent of the temperature\nor the spin-orbit coupling strength. At fixed tempera-\nture, for weak spin-orbit coupling ˜ α≪1, our expression5\nreadsM∼1−4˜α2/η2(1−cos(˜t)) [12]. In this limit,\nthe magnetization exhibits sinusoidaloscillations with an\namplitude which is proportional to 8˜ α2/η2. For strong\nspin-orbit coupling, ˜ α≫1, the magnetization becomes\nstrongly peaked near t= 2πn/ωrwherenis an integer,\nand decays exponentially, away from these points.\nThe collapse and revival of the total magnetization is\na trap effect. In a homogeneous system, the momentum\ndependent spin-orbit magnetic field will simply cause the\nspins to dephase, particles with different momenta will\nprecessatdifferentrates,andthetotalmagnetizationwill\ngo to zeroirreversiblyon a timescale set by the spin-orbit\ncoupling strength [11]. It is also important to emphasize\nthat the collapse and revival in the total magnetization\ndescribed above has a different origin from what is ob-\nserved in the experiments of Wang et al.[8]. We will\ndiscuss this in more detail later.\nTo understand the origin of the magnetization oscil-\nlations, we now turn to the dynamics of the longitudi-\nnal and transverse magnetization density following the\nramp. In a trapped geometry, from Eq. 10, we find that\nthe longitudinal magnetization density exhibits periodic\noscillations in space and in time. The temporal oscilla-\ntions have a period of t= 2π/ωr, while att=π/ωrthe\nspatial oscillations have a characteristic wave-length of\nλsh=π rtrap/(2˜α) wherertrap=/radicalbig\n/planckover2pi1/mωr.\nIn Fig. 1 we plot the magnetization density normal-\nized to the initial magnetization at the center ( m(r=\n0,t= 0)) as a function of time for the parameters\nabove. Brighter colors indicate positive magnetization\nwhile darker colors indicate negative values of mz. We\nchoose a rather weak spin-orbit coupling strength in or-\nder to enhance the wavelength of the spatial oscillations\natt=π/ωr. The transverse components of the spin\nare indicated by arrows whose length corresponds to the\nmagnitude of the spin vector in the x−yplane. At\nt= 0, all spins are pointing in the ↑direction indicated\nby the bright region in the density plot. Over time a\ntransverse component develops and a spin helix emerges.\nAtt=π/ωr, the spin oscillations reach the maximum\namplitude proportional to e−1\n2η2˜r2\n+, with a wave-length\nof Λsh/rtrap=π/2˜α.\nEnergy conserving dynamics in phase space implies\nthat the momentum of a particle is linked to its position.\nMoreover, the spin of an atom is linked to its momentum\nvia the spin-orbit coupling. The spin-orbit Hamiltonian\nshifts the minimum of the dispersion to finite momenta.\nIn a harmonically confined system, the iso-energy con-\ntours are circles in phase space, that are now shifted to\nfinite momenta due to the spin-orbit coupling. A wave-\npacket polarized in the ↑direction centered at r=p= 0\nwill follow the iso-energy contours in phase space, while\nsimultaneously rotating in spin space. At t=π/ωrthe\natomic wave-packet is centered around r= 0 in real\nspace, and in order to conserve the total energy, the dis-\ntribution will be centered around ˜ p+=±2˜αηin momen-\ntum space. As a result, atoms with opposite momenta\nprecess in opposite directions in spin space, producing aspin helix. At t= 2π/ωr, the atoms return to their orig-\ninal distribution in real and momentum space, and the\ninitial state is recovered. The spin helix has a smaller net\nmagnetization as compared to the fully polarized initial\nstate, thus explaining the oscillations in the net magne-\ntization.\nExperimentally, the spin density can be imaged in situ\nusing phase contrast imaging [36, 37]. The wave-length\nof the spin helix is set by the spin-orbit interaction which\nis determined by the wave-length of the Raman beams.\nFor the parameters used in the experiment of Cheuk et\nal., Λsh∼0.5µm [9], which may be below the experi-\nmental resolution. However, the wave-length can be in-\ncreased by decreasing the spin-orbit coupling strength.\nAt present, experiments on spin-orbit coupled cold gases\nsuffer from extremely short lifetimes ( <500ms) due to\nthe large heating rates resulting from inelastic light scat-\ntering from the Raman beams [38]. For typical trapping\npotentials, the timescale for the appearance of the spin\nhelix ist=π/ωr∼50ms, so the short lifetimes may not\nbe a major limitation in observing the spin helix.\nB.ΩR/negationslash= 0\nWe now turn to the experimentally relevantcase where\nuponsuddenlyswitchingontheRamanbeams, theatoms\nalsoexperienceaconstant, momentumindependent mag-\nnetic field, parametrized by a dimensionless parameter\n˜B= ΩR/ωr. In this limit, the problem cannot be solved\nanalytically as the Zeeman and spin-orbit terms in Eq. 5\ndonotcommute. Instead we numerically integrate the\nBoltzmann equation on a 4D grid in phase space. We\nchoose a 20×20 lattice in Randpwith a spatial res-\nolution ofδr= 2rtrap, wherertrap=/radicalbig\n/planckover2pi1/mωris the\ncharacteristic length scale of motion in the trap, and mo-\nmentum resolution of δp= 0.6 2π/Λth. The integration\nis done using a split-step method that conserves total\nparticle number and the total energy to high accuracy\nfor sufficiently small time steps.\nFor simplicity we choose a fully polarized initial state,\nwhich is stationary in the absence of interactions or spin-\norbitcoupling. Wethenconsidertwolimits, uponswitch-\ning on the Raman beams: ˜B∼˜αand˜B >˜α. The\ntwo parameters can be controlled independently as the\nmagnitude of ˜ αis set by the wave-length of the Raman\nbeams, and Ω Ris set by the laser intensity. The result-\ning dynamics of the total magnetization as well as the\nmagnetization density is plotted in Fig. 2.\nAs shown in Fig. 2, the addition of a Zeeman term\ncauses the magnetic field oscillations to decay over time.\nAt long times, the total magnetization acquires a new\nsteady state value which is smaller than 1. As the\nstrength of the Zeeman field is increased, the fully po-\nlarized initial state becomes increasingly stable, and the\ntotal magnetization remains close to 1 at long times with\nsmall oscillations.\nThe dynamics can be understood as follows: in6\n02Π4Π6Π8Π1\n0.5\n1\n0.5\n0\n1\nt/LParen11/Slash1Ωr/RParen1M/LParen1t/RParen1/Slash1N/Minus20/Minus1001020/Minus0.400.40.8\nr/Slash1rtrapS/LParen1r/RParen1/Slash1n\n/Minus20/Minus100102000.40.8\nr/Slash1rtrapS/LParen1r/RParen1/Slash1n\n/Minus20/Minus100102000.40.8\nr/Slash1rtrapS/LParen1r/RParen1/Slash1n\nFIG. 2: (Left) Time Evolution of the total magnetization\nin the system (see Eq. 9) normalized to the total parti-\ncle number for three different values of the dimensionless\nZeeman coupling ˜B= ΩR/ωr. In each case, we fix ˜ α=√\n2α/radicalbig\nm//planckover2pi1ωr= 0.25. From top to bottom: (Black) ˜B= 0;\n(Blue)˜B= 0.25; (Red) ˜B= 1. (Right) Magnetization\ndensity along the r+=x+ydirection for the same val-\nues of the spin-orbit coupling strength and Zeeman field as\nin the left figure at fixed time t=π/ωr. The magnetiza-\ntion densities are normalized to the total central density d e-\nnotedn=n(r= 0,t= 0) =/integraltext\ndp[f↑↑(p,0,0) +f↓↓(p,0,0)].\nSolid curves represent the magnetization density in the zdi-\nrection (Eq. 9), while the dashed curves indicate the mag-\nnetization density in the xdirection: mx(r,t=π/ωr) =/integraltext\ndp[f↑↓(p,r,π/ωr)+f↓↑(p,r,π/ωr)].\nthe presence of a Zeeman field, each spin precesses\nabout a new magnetic field, which is the sum of the\nspin-orbit magnetic field and the Zeeman field, and is\ntilted away from the x-y plane by an angle sin( θp) =\n˜B//radicalBig\n˜B2+(˜αη˜p+)2. As the initial state has a Gaussian\ndistribution of atoms with different momenta with spins\npointinginthe z−direction,atomswith momentagreater\nthanp>˜pcrit=˜B/˜αηwill primarilyexperience the spin-\norbit magnetic field and precess about the x−yplane,\nwhile atoms with momenta p>/planckover2pi1ΩR), the majority of the atoms still expe-\nrience the spin-orbit magnetic field, and the spin den-\nsity wave is preserved (as shown in the blue curves in\nFig. 2) for the first few oscillations. On longer timescales\nthe magnetic field affects the spin precession of even the\natoms with p >> p critand the spin density wave disap-\npears. On very long times the system settles into a new\nsteady state with a lower net magnetization.\nOn the other hand, if the Zeeman field is strong(αpth<(N−2)π.\nThus, the subsystem containing the first and the last\nspins of the chain stays in an unentangled state un-\ntil the last kick is applied. As a result the subsys-\ntem reaches a maximally entangled state at the mo-\nmentst= (N+2k−1)π, k= 0,1,..., being unen-\ntangled with the rest of the chain, in accordance with\nthe monogamy property [21] .So that, the time needed\nto completely entangle the ends of the chain linearly de-\npends on the length of the chain. In particular, at the\ninstantt= (N−1)πthe state of the chain is as follows:\n|Ψ/angbracketright=1\n2(N−1)/2N−1/productdisplay\nj=2/bracketleftBig\n|1/angbracketrightj+(−1)j|0/angbracketrightj/bracketrightBig\n[|00/angbracketright1N+|11/angbracketright1N+i(|01/angbracketright1N+|10/angbracketright1N)].\nOn the other hand, the concurrence between the spins\n1 and 2, and the spins N−1 andN, remains zero for any\nt > π.3\n0 5 10 15 20 25 3000.20.40.60.81\n(a)\n0 5 10 15 20 25 3000.20.40.60.81\n(b)t1t2\nt1t2t3t4t5\nt\nFIG. 1: (Color online) Dynamics of concurrence C(t) for the\nchain with (a) 4 spins for t1=π,t2= 2π; and (b) 7 spins\nfort1=π,t2= 2π,t3= 3π,t4= 4π,t5= 5π. Dashed line\nshowsTr(ρ2\n1N), arrows indicate when the kicked pulsed are\napplpied.\nIn Fig.(1) we show the dynamics of the concurrence\nC(t) for the chain with (a) 4 spins and π-pulses applied\nto the first 3 spins at tj=jπ,j= 1,2, and (b) 7 spins\nandπ-pulses applied to the first 6 spins at tj=jπ,\nj= 1,...5. The dashed line shows the purity of the of\nsubsystem composed by the first and the last spins of\nthe corresponding chain, Tr(ρ2\n1N). Observe that, due to\nthe Hamiltonian evolution, the concurrence and purity\nkeeps oscillating after application of the last kick.\nIn order to study the behavior of concurrence C(t)\nwhen the times tjof application of the π-pulses are not\ncommensurable, we have numerically calculated C(t) =\nC(t,t1,t2) as a function of t1andt2> t1at different\nfixed times. In Fig.2 we plot the concurrence C(t) for 4\nspins at time t= 3π(compare with Fig.1 (a)) when t1\n∈[0.1,5] andt2∈[5.1,9] . We can observe that C(t1,t2)\nhas a pronounced maximum, C= 1,att1=π, and\nt2= 2πand smoothly decreases when t1,2deviate from\nthese optimal values.\nIV. DISCUSSION\nWe have studied the dynamics a one-dimensional open\nIsing chain of Nspins assisted by phase kicks at times051015202530\n05101520253000.20.40.60.81 \nt2t1 \nFIG. 2: (Color online) The concurrence Cfor chain of 4 spins\nas a function of the moments of applied kicks at fixed time\nt= 3π.\ntj=jπ,j= 1,2,...,N−2. It is found that the ap-\nplication of N−2 instant pulses to N−1 spins leads\nto arising of transient perfect entanglement between the\nfirst and the last spins of the chain. One interesting re-\nsults is that, if the number of pulses is less than required,\nthen every pairwise concurrence would be zero. The pe-\nriodic behavior of the concurrence between the ends of\nthe chain, C1N(t) reaches the maximum value C= 1 at\nsome specific times even when the pulses are stopped be-\ning applied and the chain evolves under the Hamiltonian\n(1) only. This effect substantially enlarges possible ap-\nplications of the low-dimensional spin chains in quantum\ninformation technology.\nOn the other hand, in this particular model we may\nobserve an important example of a non-trivial effect of\nlocal transformations on the entanglement production in\nnon-linear systems.\nIt is easy to see that combining the present scheme\nwith quench one can entangle any two spins rands,\n1< r < s < N , in the chain, i.e. to model a quantum\nrouter. Really, we just have to “disconnect” the chain\n[10] from 1 to r−1 and from s+1 toNand then apply\nthe above discussed sequence of pulses.\nIt worth noting that similar local transformations in\nthe form of instant pulses were also used for another pur-\nposes: for atomic squeezing enhancement in the Dicke\nstates [22], for intensification of the entanglement in\ncontinuous-variables [23] and in two-spin [24] systems.\nThis work is partially supported by the Grant 106525\nof CONACyT (Mexico)4\n[1] M. Horodecki, P. Horodecki, and R. Horodecki, Phys.\nRev. A60, 1888 (1999).\n[2] A. Bayat and S. Bose, Phys. Rev. A 81, 012304 (2010).\n[3] L. Amico, R. Fazio, A. Osterloh and V. Vedral,\nRev.Mod.Phys. 80, 517 (2008).\n[4] S. Bose, Contemporary Physics 48, 13 (2007).\n[5] A. Osterloh, L. Amico, G. Falci, and R. Facio, Nature\n416, 608 (2002); T. J. Osborne, and M. A. Nielsen, Phys.\nRev. A66, 032110 (2002).\n[6] S. Bose, Phys. Rev. Lett. 91, 207901 (2003)\n[7] M. H. Yungand S.Bose, Phys.Rev. A 71, 032310 (2005).\n[8] L. CamposVenuti, C. DegliEspostiBoschi, and M.\nRoncaglia, Phys. Rev. Lett. 96, 247206 (2006); L. Cam-\nposVenuti, S. M. Giampaolo, F. Illuminati, and P. Za-\nnardi, Phys. Rev. A 76, 052328 (2007); A. Ferreira, J.\nM. B. L. dosSantos, Phys. Rev. A 77, 034301 (2008).\n[9] H. Wichterich and S. Bose, Phys. Rev. A 79, 060302(R)\n(2009).\n[10] A. Bayat, S. Bose, and P. Sodano, Preprint\narXiv:1007.4516 (2010).\n[11] C. Di Franco, M. Paternostro, and M. S. Kim, Phys.\nRev.A77, 020303(R) (2008).\n[12] X. Wang, A. Bayat, S. G. Schirmer, and S. Bose, Phys.\nRev. A81, 032312 (2010).\n[13] F. Galve, D. Zueco, S. Kohler, E. Lutz, P. H¨ anggi,\nPhys. Rev. A 79, 032332 (2009); F. Galve, D. Zueco,\nG. M. Reuther, S. Kohler, and P. H¨ anggi, Europ. Phys.\nJ: Special Topics 180, 237 (2009).\n[14] G. P. Berman, G. D. Doolen, G. D. Holm, andV. I. Tsifrinovich, Phys. Lett. A 193, 444 (1994);\nI. L. Chuang, N. Gershenfeld, and M. Kubinec,\nPhys. Rev. Lett. 80, 3408 (1998).\n[15] H. J. Briegel et al, Nature Physics 5, 19 (2009).\n[16] J. Fitzsimons and J. Twamley, Phys. Rev. Lett. 97,\n090502 (2006)\n[17] L. Viola, and S. Lloyd, Phys. Rev. A 58, 2733 (1998); M.\nBan, J. Mod. Opt. 45, 2315 (1998).\n[18] G. S. Uhrig, Phys. Rev. Lett. 98, 100504 (2007); K.\nKhodjasteh and D. A. Lidar, Phys. Rev. Lett. 95, 180501\n(2005); W. Yang and R. B. Liu, Phys. Rev. Lett. 101,\n180403 (2008); G. S. Uhrig, New J. Phys. 10, 083024\n(2008).\n[19] G.S. Agarwal, arXiv:0911.4158 (2009).\n[20] W. K. Wootters, Phys. Rev. Lett. 80, 2245 (1998).\n[21] V. Coffman, J. Kundu, and W. K. Wootters, Phys. Rev.\nA61, 052306 (2000); T. J. Osborne and F. Verstraete,\nPhys. Rev. Lett. 96, 220503 (2006).\n[22] D. Shindo, A. Chavez,S. M. Chumakov, and\nA. B. Klimov, J. Opt. B: Quamtum Semiclassical\nOpt.6, 34 (2004).\n[23] B. Kraus, K. Hammerer, G. Giedke, and J. I. Cirac,\nPhys. Rev. A 67, 042314 (2003).; K. Hammerer,\nK. Mølmer, E. S. Polzik, and J. I. Cirac, Phys. Rev. A\n70, 044304 (2004).\n[24] G. Burlak, I. Sainz and A.B. Klimov, Phys. Rev. A 80,\n024301 (2009)." }, { "title": "0909.3711v3.Engineering_ultralong_spin_coherence_in_two_dimensional_hole_systems_at_low_temperatures.pdf", "content": "arXiv:0909.3711v3 [cond-mat.mes-hall] 12 Feb 2010Engineering ultralong spin coherence in\ntwo-dimensional hole systems at low temperatures\nT Korn, M Kugler, M Griesbeck, R Schulz, A Wagner, M\nHirmer, C Gerl, D Schuh, W Wegscheider ‡and C Sch¨ uller\nE-mail:tobias.korn@physik.uni-regensburg.de\nInstitut f¨ ur Experimentelle und Angewandte Physik, Universit¨ at Regensburg,\nD-93040 Regensburg, Germany\nAbstract. For the realisation of scalable solid-state quantum-bit systems, sp ins\nin semiconductor quantum dots are promising candidates. A key req uirement for\nquantum logic operations is a sufficiently long coherence time of the sp in system.\nRecently, hole spins in III-V-based quantum dots were discussed a s alternatives to\nelectron spins, since the hole spin, in contrast to the electron spin, is not affected by\ncontacthyperfine interactionwith the nuclearspins. Here, we rep orta breakthroughin\nthe spin coherence times of hole ensembles, confined in so called natu ral quantum dots,\nin narrow GaAs/AlGaAs quantum wells at temperatures below 500 mK. Consistently,\ntime-resolvedFaradayrotationand resonantspin amplification tec hniquesdeliver hole-\nspincoherencetimes, whichapproachinthe lowmagneticfieldlimit value sabove70ns.\nThe optical initialisation of the hole spin polarisation, as well as the inte rconnected\nelectron and hole spin dynamics in our samples are well reproduced us ing a rate\nequation model.\nPACS numbers: 78.67.De, 78.55.Cr\nSubmitted to: New J. Phys.\n‡present address: Solid State Physics Laboratory, ETH Zurich, 80 93 Zurich, SwitzerlandEngineering ultralong spin coherence in two-dimensional h ole systems 2\n1. Introduction\nAmongthemostpromisingsystems fortherealisationofquantumco mputingdevices are\nspins in semiconductor quantum dots (QDs) [1]. Using electrostatic g ating techniques,\nthese dots may be defined within a two-dimensional electron system (2DES) by local\ndepletion. This approach has the advantage that it allows for the fa brication of\nscalable arrays of quantum bits, as it is based on high-resolution litho graphy techniques\ninstead of self-organized growth of QDs. While the spin dephasing of electrons in high-\nmobility GaAs/AlGaAs-based 2DES is extremely fast - on the order of only a few\ntens of picoseconds [2, 3, 4] in the low-excitation limit - for electron s confined in QDs,\nspin dephasing is strongly reduced and the main remaining spin dephas ing channel is\ncontact hyperfine interaction with the nuclei [5]. The latter may be suppressed by using\nelaborate spin echo techniques [6]. Several alternative material sy stems like silicon [7]\nand graphene [8] have been suggested to overcome the problem of hyperfine interaction\nof electrons and nuclei. Recently, hole spins confined in QDs have bee n considered for\nquantum computing, as long hole spin dephasing times (SDT) were obs erved in p-doped\nself-organized QDs [9, 10]. Additionally, efficient electrical tuning of t he effective hole g\nfactorinlow-dimensional structureswaspredicted[11]andexper imentally demonstrated\n[12]. In the present work we demonstrate that very long hole SDTs c an be observed in\na two-dimensional hole system (2DHS), residing in a narrow quantum well (QW) in a\nGaAs/AlGaAs-based heterostructure. We show that the hole SDT strongly depends on\nthe energy splitting between the quantized heavy-hole (HH) and ligh t-hole (LH) states\nwithin the QW, which is controlled by the QW width. Specifically, the hole S DT can\nreach values above 70 ns at low temperatures and in small magnetic fi elds, which is\nabout two orders of magnitude longer than previously reported re sults on wider GaAs\nQWs [12, 14]. In particular, these long SDTs allow us to use resonant s pin amplification\n(RSA)techniques forprecisemeasurementsoftheholespindynam ics. Theresultsshown\nhere suggest that GaAs-based 2DHS may be a viable alternative to 2 DES for quantum\ncomputing applications, which rely on electrostatically confined char ged carriers.\n2. Sample design and experimental methods\nOur samples are single-side p-modulation-doped GaAs/Al 0.3Ga0.7As QWs containing\na 2DHS with relatively low hole density (see Table 1). Previous investiga tions on\nsimilar samples with relatively wide wells (15 nm and 10 nm) [12] showed th at at\nthese low densities, the optical recombination spectra at liquid-Heliu m temperatures\nare governed by recombinations of neutral and positively-charge d excitons, i.e., bound\nexcitonic complexes, consisting of one electron and two holes. Most importantly, even\nfor the wider QWs, at very low temperatures, the resident holes be come localised in\npotential fluctuations in the plane of the QW [12]. It was shown by Syp erek et al. [14]\nand Kugler et al. [12] that localisation of the holes is crucial for the ob servation of long\nhole SDTs on the order of a few-hundred picoseconds in those samp les. Here, we reportEngineering ultralong spin coherence in two-dimensional h ole systems 3\nTable 1. Sample data. Density and mobility were determined from\nmagnetotransport measurements at 1.3 K.\nSample QW width hole density phole mobility µelectron g\n(nm) (1011cm−2) (105cm2/Vs) factor |ge|\nA 15 0.90 5.0 0 .280±0.005\nB 9 1.03 3.6 0 .133±0.01\nC 7.5 1.10 5.3 0 .106±0.01\nD 4 1.10 0.13 0 .266±0.003\non hole SDTs in narrow QWs, which are up to two orders of magnitude lo nger.\nThe structures are grown by molecular-beam epitaxy (MBE) on [001 ] substrates.\nSome characteristic properties are listed in Table 1. The table clearly shows how the\nhole mobility is significantly reduced for the thinnest sample D, most like ly due to\nthe increased influence of monolayer fluctuations at the AlGaAs/Ga As interfaces on\nthe hole wave function within the QW. For time-resolved Faraday rot ation (TRFR)\nmeasurements, the samples are first glued onto a sapphire substr ate with optically\ntransparent glue, then the semiconductor substrate is removed by grinding and selective\nwet etching. All samples contain a short-period GaAs/AlGaAs super lattice, which\nserves as an etch stop, leaving only the MBE-grown layers. The TRF R and resonant\nspin amplification (RSA) measurements are performed in an optical c ryostat with3He\ninsert, allowing for sample temperatures below 400 mK and magnetic fi elds of up to\n11.5 Tesla. A pulsed Ti-Sapphire laser system generating pulses with a length of 600 fs\nand a spectral width of 3-4 meV is used for the optical measuremen ts. The repetition\nrate of the laser system is 82 MHz. The laser pulses are split into a circ ularly-polarised\npump beam and a linearly-polarised probe beam by a beam splitter. A me chanical\ndelay line is used to create a variable time delay between pump and prob e. Both\nbeams are focussed to a diameter of about 80 µm on the sample using an achromat,\nresulting in an excitation density of about 2 Wcm−2. The center wavelength of the\nlaser system is tuned to near-resonance with the HH excitonic abso rption lines of the\nsamples. For this, photoluminescence spectra, taken using nonre sonant excitation, are\nusedtodetermine thetransitionenergiesoftheneutralandchar gedexcitons[12]. Asthe\nspectral linewidth of our laser system exceeds the energy splitting between the charged\nandneutral exciton transitions, bothareexcited by the pump lase r pulses. In theTRFR\nand RSA experiments, the circularly-polarised pump beam is generat ing electron-hole\npairs in the QW, with spins aligned parallel or antiparallel to the beam dir ection, i.e.,\nthe QW normal. Due to the spectral linewidth of the laser, typically, b oth, neutral and\npositively chargedexcitons areexcited near-resonantly. IntheT RFRmeasurements, the\nspinpolarisationcreatedperpendiculartothesampleplanebythepu mpbeam, isprobed\nby the time-delayed probe beam via the Faraday effect: the axis of lin ear polarisation\nof the probe beam is tilted by a small angle [13], which is proportional to the out-of-\nplane component of the spin polarisation. This small angle is detected using an opticalEngineering ultralong spin coherence in two-dimensional h ole systems 4\nbridge. A lock-in scheme is used to increase sensitivity. In the RSA me asurements, the\nFaraday rotation angle is measured for a fixed time delay as a functio n of an applied\nin-plane magnetic field. In our investigations, we exploit the strengt hs of both methods\nin order to explore the limits of hole spin dynamics in GaAs-based hole sy stems: In the\nTRFR experiments, one measures in the time domain via a pump-probe scheme, and\nthe SDT as well as photocarrier lifetimes can be extracted from the measurements. The\ndisadvantage of this method, however, is the limitation of the acces sible time range by\nthe travel length of the optical delay line. In our setup, this limits th e time range to\nabout 2 ns. This hinders an accurate determination of SDTs, which a re significantly\nlonger than the measurement range. The TRFR method encounter s severe problems, if\nthe SDT is even longer than the time interval between two subseque nt laser pulses in\nthe pulse train of the mode-locked laser (in our case about 12 ns). F ortunately, with\nthe RSA technique [15], one can overcome these problems, since her e one works with a\nfixed time delay between pump and probe pulses. The method is based on the resonant\namplification of the Faraday signal, when integer multiples of the prec ession period of\nthe spins in an inplane magnetic field coincide with the inverse laser repe tition rate, i.e.,\nthe time interval between subsequent pulses. However, here the extraction of the SDT\nfrom the experimental data is more involved, since it is hidden in the line widths of the\nRSA maxima. As we will show below, in particular in our case, where elect ron and hole\nrecombination and spin dynamics are interconnected, this is rather complex.\n3. Quantum-well width dependence of hole spin-dephasing ti me\nWe start our discussion by presenting TRFR experiments on our sam ples as a function\nof an in-plane magnetic field. As discussed above, in these experimen ts we are detecting\nthe charge carrier and spin dynamics of photoexcited electrons an d holes within a time\nrange<2 ns. This will allow us to gain important information on the interconnec ted\nelectron and hole photocarrier and spin dynamics. Figure 1(a) show s a series of TRFR\ntraces measured on sample D, the sample with the thinnest QW, at 1.2 K at three\ndifferent magnetic fields, applied in the sample plane. The trace at B= 0 shows a\nsingle exponential decay of the Faraday signal after pulsed excita tion with a decay time\nofτR= 70 ps. Each of the other two traces, however, exhibits the supe rposition of\ntwo damped oscillations with markedly different frequencies and damp ing constants: A\nfast oscillation, which is observable within about the decay time of the B= 0 trace,\nonly, and a slower oscillation, which, in the B= 2 T trace, exceeds the measurement\nrange by far. The sum of two damped cosine functions is fit to the da ta in order to\nextract the precession frequencies and decay constants. As a r esult, we can identify\nthe high-frequency oscillations, which decay significantly faster, a s the precession of\nphotogeneratedelectrons, andthelow-frequency oscillationsas theprecession ofresident\nholes within the sample: In figure 1(b), the dispersions of the two pr ecession frequencies\nare plotted, and the electron and hole gfactors are extracted from the data. The\nelectron gfactor|ge|= 0.266 is in good agreement with values measured for QWs ofEngineering ultralong spin coherence in two-dimensional h ole systems 5\n/s48/s46/s48 /s48/s46/s53 /s49/s46/s48 /s49/s46/s53 /s48 /s50 /s52 /s54 /s56 /s49/s48/s48/s53/s49/s48/s49/s53/s50/s48/s50/s53/s51/s48/s51/s53/s52/s48\n/s66/s61/s56/s32/s84\n/s66/s61/s50/s32/s84\n/s32/s32/s70/s97/s114/s97/s100/s97/s121/s32/s115/s105/s103/s110/s97/s108/s32/s40/s110/s111/s114/s109/s97/s108/s105/s122/s101/s100/s41\n/s116/s105/s109/s101/s32/s116/s32/s40/s110/s115/s41/s66/s61/s48/s32/s84/s40/s97/s41\n/s84/s61/s49/s46/s50/s32/s75/s40/s98/s41\n/s32/s32/s70/s114/s101/s113/s117/s101/s110/s99/s121/s32/s40/s71/s72/s122/s41\n/s77/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32/s40/s84/s41/s124/s103\n/s104/s124/s61/s48/s46/s48/s54/s53/s53/s124/s103\n/s101/s124/s61/s48/s46/s50/s54/s54\nFigure 1. TRFR for different in-plane magnetic fields and disp ersion of\nprecession frequencies. (a) TRFR traces for sample D, measured at 1.2 K for\ndifferent magnetic fields. (b) Dispersion of electron (black dots) an d hole (red stars)\nspin precession determined for sample D from TRFR measurements. The solid lines\nare linear fits to the data.\nsimilar widths [16, 17], while the hole gfactor|gh|= 0.0655 is close to zero, as was\ntheoretically predicted for GaAs-based QWs grown along the [001] d irection [18] and\nexperimentally observed [12, 14, 19].\nWe note that the decay constant of the electron spin precession, τe= 75±5 ps,\nremains almost constant throughout the investigated magnetic fie ld range. This may be\nexplained as follows: as our samples are p-doped, the optically orient ed electron spins\ncan only be observed during the photocarrier lifetime. The measure d decay constant\nτe= 75±5 ps therefore corresponds to the photocarrier lifetime in sample D ,τR, as\nthe electron SDT is typically longer. In stark contrast, the hole SDT decreases strongly\nwith increasing magnetic field, as can be seen quite drastically from th eB= 2 T and\nB= 8 T traces in figure 1(a). However, at this stage the important qu estion arises,\nwhy the trace at B= 0 in figure 1(a) obviously does not show a long lasting hole spin\nsignal but a spin signal, which is governed by the photocarrier lifetime , only? We will\naddress this question further below in the next section.\nTo investigate the strong magnetic field dependence of the hole SDT in more detail,\nwe study the hole spin dynamics as a function of the applied in-plane ma gnetic field for\nsamples with different QW widths. Figure 2(a) shows TRFR traces for all 4 samples,\nmeasured at 1.2 K with an applied in-plane magnetic field of 6 T. The trac es haveEngineering ultralong spin coherence in two-dimensional h ole systems 6\n0.0 0.2 0.4 0.6 0.8 1.0 0 2 4 6 8 100.1110\nSample D\nSample C\nSample B\n Faraday signal (norm.)\ntime t (ns)(a)\nSample AT=1.2 K\n D\n C\n B\n A(b)\n Hole SDT (ns)\nMagnetic field (T)Sample\nFigure 2. TRFR for different well widths and magnetic-field de pendence\nof SDTs. (a) TRFR traces for samples A, B, C and D at 1.2 K. A 6 Tesla in-plane\nmagnetic field was applied during the measurements. (b) Hole spin dep hasing times\nfor samples A, B, C and D as a function of magnetic field. The lines repr esent fits of\na 1/B dependence.\nbeen scaled and vertically shifted in order to allow easy comparison of the hole spin\ndynamics. In all traces, again, a fast, rapidly decaying electron sp in precession can be\nobserved within the first ∼100 ps. After photocarrier recombination, only larger-period\nhole spin precession is visible in the traces. Even though the applied in- plane magnetic\nfield was fixed at 6 T, in all measurements shown in figure 2(a), it can b e seen clearly\nthat both, the electron and the hole spin precession frequencies, are different for all 4\nsamples. For electrons, it is well-known that the gfactor in a QW depends on the QW\nwidth [16, 17]: due to the nonparabolicity of the conduction band in Ga As, the electron\ng factor changes depending on the electron energy above the con duction band edge, and\nin good approximation, the QW confinement energy of an electron will lead to a similar\nchange of the g factor. Additionally, as the electron wave function has a nonvanishing\namplitude within the QW barriers for thin GaAs QWs, the different elect rongfactors\nof pure GaAs and the barrier material are admixed. The values of th e electron gfactors\nextracted from TRFR measurements on all four samples are listed in Table 1. We note\nthat the sign of the g factor cannot be directly determined from TR FR measurements,\nhowever, by comparing our measured values to literature data [16, 17], we conclude that\nthe electron g factor for samples A-C is negative, while it is positive fo r sample D.\nOn the other hand, the different hole gfactors observed in the TRFR traces stemEngineering ultralong spin coherence in two-dimensional h ole systems 7\nfrom the strong anisotropy of the hole gtensor, which was predicted by Winkler et al.\n[18] for [001]-grown GaAs QWs. The effective hole gfactor,g∗\nh, measured in TRFR is\ngiven by the geometric sum of the in-plane ( g⊥) and out-of-plane ( g||) components of\nthe hole gtensor:\ng∗\nh=/radicalBig\ng2\n⊥cos2α+g2\n||sin2α. (1)\nHere,αis the tilt angle of the magnetic field with respect to the QW plane. While\nthe in-plane component of the hole gfactor is close to zero, g⊥∼0, the out-of-plane\ncomponent is g||∼ −0.7 for bulk GaAs [20]. Therefore, even small tilt angles αresult\nin markedly different g∗\nh.\nIn figure2(b), thehole SDTs forall 4 samples areshown asa functio n of theapplied\nmagnetic field. Two effects are clearly visible here:\n(i) For all samples, the hole SDT decreases as the magnetic field is incr eased, following\napproximately a 1 /Bdependence.\n(ii) The hole SDT increases systematically as the QW width is reduced. I t is smallest\nfor sample A with the widest QW and largest for sample D with the thinne st QW.\nThe decrease of the hole SDT with magnetic field is a well-known phenom enon. It\nis believed to be caused by the inhomogeneity of the hole gfactor, ∆ g∗\nh, which leads to\na dephasing of the hole spin polarisation due to different precession f requencies. The\ndephasing rate due to this inhomogeneity is proportional to the app lied magnetic field.\nTherefore, the hole spin dephasing time observed in the experiment , which is the spin\ndephasing timeofaninhomogeneously broadenedensemble, T∗\n2, isinfirstapproximation\ngiven by [21]\nT∗\n2=/parenleftbigg1\nT2+∆g∗\nhµBB\n¯h/parenrightbigg−1\n, (2)\nif ∆g∗\nhis considered as the only source of inhomogeneity. Here, T2is the hole spin\ndephasing time in the absence of inhomogeneous broadening.\nIn order to understand the influence of the QW width on the hole SDT , we have\nto think about the main mechanisms for hole spin dephasing. The HH an d LH states\nhave different angular momenta, transitions between these state s will therefore destroy\nhole spin orientation. In bulk GaAs, where HH and LH valence bands ar e degenerate at\nk= 0, any momentum scattering may lead to a transition between HH an d LH states,\nwhich leads to hole SDTs on the order of the momentum scattering tim e [22]. In QWs,\nthis degeneracy is lifted. However, for k >0, the valence bands have a mixed HH/LH\ncharacter [23], which may also lead to rapid hole spin dephasing due to s cattering\n[24]. At low temperatures, resident holes in QWs may become localised a t potential\nfluctuations within the QW, arising from QW thickness fluctuations du e to monolayer\nsteps at the interfaces, as well as from the granular distribution o f the remote donors.\nLocalisationsignificantlyreduces theholequasimomentum, keeping r esident holesinHH\nstates with k∼0. However, even for k∼0, there is a finite admixture of the LH states\nto the first HH subband [25]. The significant increase in hole SDT with de creasing QWEngineering ultralong spin coherence in two-dimensional h ole systems 8\nwidth can therefore be attributed to an increased HH/LH splitting, which reduces the\nLH contribution to the HH ground state [26]. In a first approximatio n, if we consider\ninfinitely high square-well potentials, this energy splitting ∆ Eis proportional to 1 over\nthe well-width Lsquared:\n∆E∼/parenleftbigg1\nmLH−1\nmHH/parenrightbigg¯h2π2\n2L2. (3)\nFor narrower QWs of finite potential height, however, the hole wav e function penetrates\nstrongly into the barrier, and the energy splitting is reduced again. For a QW width of\n4 nm, the maximum HH/LH splitting was theoretically predicted and exp erimentally\nobserved [27]. As figure 2(b) shows, we indeed observe the longest hole SDT in sample\nD with a well width of 4 nm. This suggests that the maximum HH/LH splitt ing in\nthis sample is responsible for the long SDT. We will discuss the possible a lternative\nmechanisms, which might govern hole spin dephasing in the last section . Before, we will\nin the following explore the limits of hole spin dephasing in our samples, st arting with\nthe optical initialisation process of hole spin polarisation in the next se ction.\n4. Buildup of a resident hole spin polarisation\nIn this section, we present experimental results and simulation dat a concerning the\ninitialization process of the hole spin polarisation, which was described by Syperek et\nal. [14]: excitation of the sample with circularly-polarised light will creat e spin-polarised\nelectron-hole pairs according to the optical selection rules. If the re is no significant\nelectron spin dephasing during the photocarrier lifetime, the optica lly oriented electrons\nwill recombine with holes which have matching spin orientation. Theref ore, no spin\npolarisation will remain within the sample after photocarrier recombin ation. This\nprocess is sketched schematically in the left panel of figure 3(c).\nIn an applied in-plane magnetic field, however, electrons and holes pr ecess with\ndifferent precession frequencies due to their strongly different gfactors. Therefore,\nduring their photocarrier lifetime, the electron spins will recombine w ith holes with\narbitrary spin orientation. A part of the optically oriented holes may therefore remain\nwithin the sample after photocarrier recombination, as depicted in t he right panel of\nfigure 3(c).\nIn order to gain deeper insight into the combined dynamics of electro n and hole\nspins, and to analyse our experimental results quantitatively, we s et up a model in the\nfollowing. The combined dynamics of the electron and hole spins can be described via\ncoupled differential equations for the electron ( e) and hole ( h) spin polarisation vectors:\nde\ndt=−e\nτR+geµB\n¯h(B×e) (4)\ndh\ndt=−h\nτh+ghµB\n¯h(B×h)+ezz\nτR(5)\nIn our model we assume the following: the electron spin polarisation is reduced at a\nrate, given by thephotocarrier recombination andprecesses abo utthe in-planemagneticEngineering ultralong spin coherence in two-dimensional h ole systems 9\n/s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48\n/s32/s32/s70/s97/s114/s97/s100/s97/s121/s32/s115/s105/s103/s110/s97/s108/s32/s40/s97/s114/s98/s46/s32/s117/s110/s105/s116/s115/s41\n/s116/s105/s109/s101/s32/s40/s110/s115/s41/s32/s32/s66/s61/s54/s32/s84\n/s32/s32/s66/s61/s51/s32/s84\n/s32/s32/s66/s61/s48/s32/s84\n/s102/s105/s101/s108/s100/s32/s115/s119/s101/s101/s112/s84/s61/s49/s46/s50/s32/s75\n/s48 /s50 /s52 /s54/s48/s49/s40/s98/s41/s32\n/s32/s72/s111/s108/s101/s32/s115/s112/s105/s110/s32/s115/s105/s103/s110/s97/s108/s32/s64/s32/s48/s46/s52/s32/s110/s115/s32/s40/s110/s111/s114/s109/s46/s41/s32\n/s77/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32/s40/s84/s101/s115/s108/s97/s41/s32/s70/s97/s114/s97/s100/s97/s121/s32/s115/s105/s103/s110/s97/s108\n/s32/s82/s97/s116/s101/s32/s101/s113/s110/s46/s32/s109/s111/s100/s101/s108/s40/s97/s41\nh+e-Optically□oriented\nelectron-hole□pairsResident\nholesNo□magnetic□field In-plane□magnetic□field\nh+e-\nh+e-\nh+e-recombinationprecessiontime(c)\nFigure 3. TRFR for different magnetic fields and buildup of hol e spin\npolarisation. (a) TRFR traces for sample C taken at 1.2 K for different magnetic\nfields. The arrow indicates the delay position for the magnetic field sw eep shown in\n(b). (b) Kerr signal (black dots) as a function of magnetic field tak en at a fixed time\ndelay (400 ps) between pump and probe pulses. The orange line show s the buildup\nof the hole spin polarisation as a function of magnetic field as calculate d by the rate\nequations model. (c) Diagram of the combined spin and recombination dynamics with\n(right panel) and without an applied magnetic field (left panel).Engineering ultralong spin coherence in two-dimensional h ole systems 10\n-1.0 -0.5 0.0 0.5 1.0-0.1 0.0 0.1 0.2 0.3 0.4\n0.1 1 1011010020x\nSimulationT=0.4 KT=1.2 K\n RSA signal (arb. units)\nMagnetic field (T)T=4.5 K(a) 0.4 K Experiment\n Simulation\n RSA signal (arb. units)\nMagnetic field (T)(b)n = 1234\n 0.4 K\n 1.2 K\n 4.5 KTRFR data\n Hole SDT (ns)\nMagnetic field (T)(c)RSA data\nFigure 4. RSA measurements for different temperatures and ma gnetic field\ndependence of hole SDT for different temperatures. (a) RSA traces for sample\nD, measured at different temperatures, compared to simulation da ta. (b) RSA trace\nmeasured at 400 mK (open circles) compared to simulation data (ora nge line). (c)\nHole SDT, determined from RSA data (left of the dotted line and TRFR data (right\nof the dotted line), for different temperatures as a function of ma gnetic field.\nfield vector B. Electron spin dephasing has been neglected, as the electron SDT is\nexpected to be significantly longer than the photocarrier recombin ation time τR, which\nis a reasonable assumption. The hole spin polarisation is reduced at a r ate given by the\nhole SDT τhand precesses about the in-plane magnetic field vector B. The last term in\nthe second equation describes the change of hole spin polarisation d ue to recombination\nof spin-polarised electrons with holes with matching spin orientation. Here,ezis thez\ncomponent of the electron spin polarisation, and zis the unit vector along the growth\ndirection. Asimilar approachwasused by Yugovaet al. [28] todescrib e theinitialization\nof a resident electronspin polarisation in n-doped QWs. However, we need to include,\nboth, electron and hole spin precession (second term in equation (5 )) in our differential\nequations to correctly model the resonant spin amplification spect ra we observe, as\ndescribed in the section below. In order to test the validity of this mo del, we compare it\nto experimental results. Figure 3(a) shows TRFR traces of sample C, taken at different\napplied in-plane fields. The sample has been carefully aligned so that th e magnetic field\nis applied along the QW plane, i.e., α∼0, resulting in a very low effective hole gfactor\ng∗\nh. For zero magnetic field, the TRFR signal decays monoexponentially with a decay\nconstant of τR= 48 ps, which reflects the photocarrier recombination time in sample C.\nFor larger magnetic fields, pronounced electron spin precession ca n be observed during\nτR, and a significant nonzero TRFR signal remains after photocarrier recombination.Engineering ultralong spin coherence in two-dimensional h ole systems 11\nThis signal is due to the buildup of a resident hole spin polarisation, as e xplained above.\nIts amplitude increases as the magnetic field is increased. Due to the very small hole\ngfactor (|g∗\nh|<0.005), no hole spin precession is observed in the time range shown\nin figure 3(a). This allows us to study the buildup of the hole spin polaris ation in\nmore detail using the following experimental technique: the TRFR sig nal is recorded\nfor a fixed time delay between pump and probe. The time delay is chose n such that\nphotocarrier recombination is complete, thereby leaving a TRFR sign al from resident\nholes, only. The magnetic field is ramped up from zero in small incremen ts. Figure 3(b)\nshows such a measurement for a time delay of ∆ t= 400 ps. It is clearly visible how\nthe resident hole spin polarisation increases from zero to a maximum v alue, at which\nit saturates. This buildup can easily be modeled by the differential equ ation system\ndescribed above. For this, we used the parameters extracted fr om the experiment:\n|ge|= 0.118,τR= 48 ps, g∗\nh= 0. Hole spin dephasing was neglected to model the\ndataset. The resulting hole spin polarisation as a function of magnet ic field is shown as\nan orange solid line in figure 3(b). It is in excellent agreement with the e xperimental\ndata. These measurements clearly demonstrate the crucial role o f an applied in-plane\nmagnetic field for transferring spin polarisation from optically orient ed photocarriers to\nresident holes.\n5. Resonant spin amplification measurements of hole spin dyn amics\nTo explore the hole SDT in very small applied magnetic fields, and to app roach the\nultimate limit of hole SDT in our samples, we employ the RSA technique [15 ], which\nhasbeenusedinrecentyears, e.g.,tostudyelectronspindynamics inn-dopedbulkGaAs\n[29] and 2DES in CdTe-based QWs [30]. RSA is based on the constructiv e interference\nof spin polarisations created by subsequent laser pulses in a time-re solved Faraday or\nKerr rotation measurement. The Faraday/Kerr signal is measure d as a function of the\nin-plane magnetic field for a fixed time delay ∆ t, typically before the arrival of a pump\npulse. If the optically oriented spins precess by an integer multiple of 2πwithin the\ntime delay between two pump pulses, a maximum in the RSA signal is obse rved. The\nmaxima are typically periodic in B. The SDT can be determined from the half-width\nof the maxima, and their spacing yields the gfactor. In systems where the buildup\nof a resident spin polarisation is governed by the interplay of electro n and hole spin\ndynamics, however, a more complex shape of the RSA signal can be e xpected, as will\nbe demonstrated now.\nFigure 4(a) shows a series of RSA traces, measured on sample D for different\ntemperatures, compared to a numeric simulation, using the coupled differential equation\nmodel introduced above. The unusual ’butterfly’ shape [28] of t he RSA signals is\nclearly visible for the traces measured at 1.2 K and 0.4 K. It is well-repr oduced by\nthe simulation data. The RSA signal at 4.5 K is about 20 times weaker th an the\nlower-temperature signals. In figure 4(b), a closeup of the first f ew RSA maxima\nmeasured at 0.4 K is compared to simulation data. There is no RSA maxim um inEngineering ultralong spin coherence in two-dimensional h ole systems 12\nthe measurement and the simulation for B= 0, in contrast to typical RSA curves\nmeasured, e.g., in n-doped GaAs bulk. This is due to the peculiar trans fer process\nthat leads to a resident hole spin polarisation. As demonstrated abo ve, a finite in-plane\nmagneticfieldisnecessarytocreatearesident holespinpolarisation . Thefirstmaximum\nat finite field (numbered as n= 1) has a distinct shape resembling the derivative of a\nLorentzian, while subsequent maxima resemble slightly asymmetrical Lorentzian curves.\nThe amplitude of the maxima first increases with n, then decreases again for n >5 (see\nfigure 4(a)), while the FWHM of the maxima increases. These featur es are clearly\nreproduced in the simulation. They can be explained as follows: for sm all magnetic\nfields, there is only a partial transfer of spin polarisation to the res ident holes, a process\nwhich saturates in our experiment for B∼0.5 T. For larger magnetic fields, the RSA\nFWHM increases due to inhomogeneous broadening, and the amplitud es of the RSA\nmaxima decrease accordingly, until the RSA signal drops to the nois e level at about\n1.6 T in the low-temperature measurements. Interestingly, direct comparison of the\nRSA traces measured at 1.2 K and 4.5 K reveals that the effective hole g factor |g∗\nh|\ndecreases from 0.067 to 0.050 as the temperature is increased, in g ood agreement with\nprevious observations [14].\nFigure 4(c) shows the hole spin lifetime insample D for different temper atures. The\ndata for low magnetic fields (below 1 T) have been determined from RS A data, the data\nfor higher magnetic fields (above 2 T) were determined from TRFR da ta. Interestingly,\nwhile at 4.5 K, the hole SDT saturates at about 2.5 ns, even in low magne tic fields,\nit follows a clear 1 /B-like dependence down to very small magnetic fields in the data\nmeasured at lower temperatures, yielding values of T∗\n2= 74±15 ns at 0.4 K and\nT∗\n2= 61±11 ns at 1.2 K (determined from the FWHM of the second RSA maximum a t\nB∼0.2 T) [36]. The power of the RSA technique is evident if one compares th e hole\nSDT measured at higher fields: in the magnetic field range where TRFR measurements\nyield precise results of the hole SDT, very little difference is observed in the values of\nthe hole SDT in the temperature range from 0.4 K to 4.5 K. From the ma gnetic field\ndependence oftheholeSDT, wemayinferthatthe T2timeoftheholespinisabove80ns\nat temperatures below 500 mK. By fitting equation 2 to the magnetic field dependence,\nwe determine the g factor inhomogeneity ∆ g∗\nh= 0.003±0.0002.\nFurthermore, the RSA measurements allow for a precise determina tion of the\neffective hole gfactor, which is given by g∗\nh= 2πfrep¯h/(µB∆B). Here, frepis the laser\npulse repetition frequency, and ∆ Bis the magnetic field spacing between two adjacent\nRSA maxima. Figure 5 shows RSA traces measured at 1.2 K for differen t angles α\nof the magnetic field with respect to the QW plane [35]. The geometry is sketched in\nthe figure. The spacing of the maxima is significantly reduced as the m agnetic field\nacquires an out-of-plane component. The effective hole gfactor|g∗\nh|is extracted from\nthis spacing, the results are shown in the inset. The increase of the effective hole gfactor\nwith the magnetic field angle is due to an admixture of the out-of-plan e component of\nthe hole gfactorg||, which is typically far larger than the in-plane hole gfactorg⊥,\nas described by equation 1. By fitting the results with this equation a s indicated byEngineering ultralong spin coherence in two-dimensional h ole systems 13\n0.0 0.1 0.2 0.3 0.4 0.5-10 -5 0 5 100.050.100.15\nα\n RSA signal (norm.)\nMagnetic field (T)0o4.5o\n3oB\nT=1.2 K\n eff. hole g factorTilt angle α (deg.)\nFigure 5. RSA measurements for different tilt angles of the ma gnetic field\nand effective hole g factor dependence on tilt angle. RSA traces for sample D,\nmeasured at 1.2 K, for different tilt angles of the sample with respect to the external\nmagnetic field. The inset shows the effective hole g factor |g∗\nh|determined from the\nspacing of the RSA maxima (black dots). The solid red line represents a fit of the\nresults using equation 1.\nthe solid red line in the inset, both components of the hole g factor at 1.2 K can be\ndetermined with high accuracy: |g⊥|= 0.059±0.003,|g|||= 0.89±0.03.\nFinally, we will discuss the possible mechanisms, which might limit the hole S DT\nin our experiments. At low temperatures and in low magnetic fields, wh ere thegfactor\ninhomogeneity may be neglected, in principle several mechanisms migh t be thought\nof to pose the ultimate limit of the hole SDT. In the following, we will discu ss their\ndependence on QW thickness in the light of our experiments, and iden tify the most\nimportant mechanism.\nThermal activation of holes from localized states: The thermal activation rate is\nproportional to the Boltzmann factor exp/parenleftBig\n(−ELoc)/(kBT)/parenrightBig\n. For localization of holes at\nQW monolayer thickness fluctuations, the localization energy increa ses drastically with\nareductionoftheQWwidth[25]. Fora15nmwideQW,itislessthan1meV, whilefora\n4 nm wide QW, it is about 10 meV. In both cases, the localization energy is significantly\nlargerthanthethermalenergy, evenatthehighestmeasuremen ttemperatureusedinour\nexperiments: Eth(4.2 K)=0.36 meV. This indicates that thermal activation of holes outEngineering ultralong spin coherence in two-dimensional h ole systems 14\nof localized states is unlikely to affect hole spin dynamics at liquid-helium t emperatures\nand below.\nDipole-Dipole interaction of holes with nuclei: Even though the p-like symmetry of\nvalence band states does not allow for a contacthyperfine interaction between holes\nand nuclei, a coupling is possible via dipolar interactions [31, 10]. The nuc lei act as\nslowly varying random magnetic fields, which lead to both, single spin de coherence\nand ensemble spin dephasing. The variance of the nuclear magnetic fi eld is inversely\nproportional to thenumber of nuclei that interact with the hole co nfined in the quantum\ndot, which maybe estimated fromthe dot size. Wenotethat thetyp ical size of quantum\ndots which arise from monolayer thickness fluctuations in high-qualit y GaAs QWs (TF-\nQDs) is on the order of (100 nm)2, significantly larger than that of self-assembled InAs\ndots. One may estimate that a hole interacts with about 106nuclei in TF-QDs or\nelectrostatically controlled QDs [6], whereas self-assembled dots ty pically contain only\nabout 5·104nuclei [10]. Additionally, the random magnetic fields associated with th e\nnuclei may be suppressed very efficiently by a small, external magne tic field applied\nwithin the sample plane [31]. It is shown above that the application of an in-plane\nmagnetic field is necessary for the transfer of spin polarisation of o ptically oriented\ncarriers to resident holes. Therefore, we may neglect hole spin dec oherence due to\ninteractionwithnucleiasthedominantprocess, limitingtheholeSDTin ourexperiment.\nFinite admixture of LH states to HH states: Even localized holes have a finite\nquasimomentum, which is given by the hole temperature. In an analog on to the Elliott-\nYafet [32, 33] mechanism, which has been studied in detail for electr ons, any momentum\nscattering may therefore destroy hole spin orientation via hole spin flip due to the slight\nadmixtureofLHstatestoHHstates. Thisadmixtureisontheorder of2percent forthin\nQWs [25]. It increases significantly for wider QWs due to the decreasin g HH/LH energy\nsplitting, increasing the probability of a hole spin flip during momentum s cattering [34].\nThespin dephasing rateisdirectly proportionaltothemomentum sc attering rate, which\nin turn increases with the sample temperature.\nSummarizing this part: We believe that the increased HH/LH splitting is\nresponsible for the increased hole SDT with decreasing QW width in our experiments.\nOn the other hand, the still finite LH admixture to the HH ground sta te and, hence,\nthe possibility of spinflip scattering of Elliott-Yafet type sets the limit for hole spin\ndephasing.\nConclusions\nIn conclusion, we have investigated hole spin dynamics in two-dimensio nal hole systems\nembedded in quantum wells of different width. Due to the increased en ergy splitting\nbetween heavy- and light-hole states, the hole spin dephasing time in creases drastically\nin narrow quantum wells. The inhomogeneity ∆ g∗\nhof the hole gfactor leads to aEngineering ultralong spin coherence in two-dimensional h ole systems 15\ntypical 1/B-like dependence of the hole SDT. In order to transfer spin polarisa tion from\nthe optically oriented photocarriers to the 2DHS, a finite magnetic fi eld is necessary.\nThis also leads to a peculiar shape of resonant spin amplification signals measured on\n2DHS. From these RSA signals, the hole SDT in low magnetic fields can be extracted.\nIt is on the order of 80 ns, one order of magnitude larger than for e lectrons in\nquantum dots of similar dimensions defined in 2DES by external gates [6]. This makes\n2DHS an interesting material system for scalable quantum computin g devices based\non electrostatically confined charge carriers. Additionally, RSA mea surements in tilted\nmagnetic fields allow us to accurately determine both, the in-plane an d out-of plane\ncomponents of the hole g factor.\nAcknowledgments\nThe authors would like to thank E.L. Ivchenko, M.M. Glazov and M.W. Wu for fruitful\ndiscussion. Financial support by the DFG via SPP 1285 and SFB 689 is g ratefully\nacknowledged.\nReferences\n[1] Awschalom D D, Loss D and Samarth N 2002 Semiconductor Spintronics and Quantum\nComputation (Berlin: Springer)\n[2] Brand M A, Malinowski A, Karimov O Z, Marsden P A, Harley R T, Shield s A J, Sanvitto D,\nRitchie D A and Simmons M Y 2002 Phys. Rev. Lett. 89236601\n[3] Stich D, Zhou J, Korn T, Schulz R, Schuh D, Wegscheider W, Wu M W a nd Sch¨ uller C 2007 Phys.\nRev. Lett. 98176401\n[4] Stich D, Zhou J, Korn T, Schulz R, Schuh D, Wegscheider W, Wu M W a nd Sch¨ uller C 2007 Phys.\nRev.B76073309\n[5] Khaetskii A V, Loss D and Glazman L 2002 Phys. Rev. Lett. 88186802\n[6] Petta J R, Johnson A C, Taylor J M, Laird E A, Yacoby A, Lukin M D, M arcus C M, Hanson M\nP and Gossard A C 2005 Science 3092180\n[7] Eriksson M A, Friesen M, Coppersmith S N, Joynt R, Klein L J, Slinker K, Tahan C, Mooney P\nM, Chu J O and Koester S J 2004 Quantum Information Processing 3133\n[8] Trauzettel B, Bulaev D V, Loss D and Burkard G 2007 Nature Phy s.3192\n[9] Heiss D, Schaeck S, Huebl H, Bichler M, Abstreiter G, Finley J, Bula ev D V and Loss D 2007 Phys.\nRev.B76241306\n[10] Eble B, Testelin C, Desfonds P, Bernardot F, Balocchi A, Amand T, Miard A, Lemaˆ ıtre A, Marie\nX and Chamarro M 2009 Phys. Rev. Lett. 102146601\n[11] Andlauer T and Vogl P 2009 Phys. Rev. B79045307\n[12] Kugler M, Andlauer T, Korn T, Wagner A, Fehringer S, Schulz R, K ubova M, Gerl C, Schuh D,\nWegscheider W, Vogl P and Sch¨ uller C 2009 Phys. Rev. B80035325\n[13] forthesamplesinvestigatedhere, themaximumTRFRsignalcor respondstoatilt angleof70 µRad\n[14] Syperek M, Yakovlev D R, Greilich A, Misiewicz J, Bayer M, Reuter D and Wieck A D 2007 Phys.\nRev. Lett. 99187401\n[15] Kikkawa J M and Awschalom D D, Phys. Rev. Lett. 801998 4313\n[16] Snelling M J, Flinn G P, Plaut A S, Harley R T, Tropper A C, Eccleston R and Phillips C C 1991\nPhys. Rev. B4411345\n[17] Yugova I A, Greilich A, Yakovlev D R,Kiselev A A,Bayer M,Petrov V V,Do lgikh Yu K,Reuter D\nand Wieck A D 2007 Phys. Rev. B75245302Engineering ultralong spin coherence in two-dimensional h ole systems 16\n[18] Winkler R, Papadakis S J, De Poortere E P and Shayegan M 2001 Phys. Rev. Lett. 854574\n[19] Marie X, Amand T, Le Jeune P, Paillard M, Renucci P, Golub L E, Dym nikov V D and Ivchenko\nE L 1991 Phys. Rev. B605811\n[20] Landolt-B¨ ornstein 1987 Semiconductors: Intrinsic Properties of Group IV Elements and III-V, II-\nVI and I-VII Compounds New Series Group III Vol. 22Pt. A ed. O. Madelung (Berlin: Springer)\n[21] YakovlevD R and BayerM 2008 Coherent Spin Dynamics of carriers, chapter 6 in Spin Physic s in\nSemiconductors Springer Series in Solid-State Sciences 157ed. Dyakonov M I (Berlin: Springer)\n[22] Hilton D J and Tang C L 2002 Phys. Rev. Lett. 89146601\n[23] Pfalz S, Winkler R, Nowitzki T, Reuter D, Wieck A D, H¨ agele D and O estreich M 2005 Phys. Rev.\nB71165305\n[24] Damen T C, Vi˜ na L, Cunningham J E, Shah J E and Sham L J 1991 Phys. Rev. Lett. 673432\n[25] Luo J-W, Bester G and Zunger A 2009 Phys. Rev. B79125329\n[26] Winkler R 2003 Spin-orbit Coupling Effects in 2D Electron and Hole Systems (Berlin: Springer)\n[27] El Khalifi Y, Gil B, Mathieu H, Fukunaga T and Nakashima H 1989 Phys. Rev. B3913533\n[28] Yugova I A, Sokolova A A, Yakovlev D R, Greilich A, Reuter D, Wieck A D and Bayer M 2009\nPhys. Rev. Lett. 102167402\n[29] Kikkawa J J and Awschalom D D 1999 Nature 397, 139\n[30] Zhukov E A, Yakovlev D R, Bayer M, Karczewski G, Wojtowicz T a nd Kossut J 2006 Phys. Status\nSolidi(b)243878\n[31] Fischer J, Coish W A, Bulaev D V and Loss D 2008 Phys. Rev. B78155329\n[32] Elliott R J 1954 Phys. Rev. 96266\n[33] Yafet Y 1963 Solid State Phys. 141\n[34] L¨ u C, Cheng J L and Wu M W 2005 Phys. Rev. B71075308\n[35] The sample tilt angle is determined by measuring the angle of the pu mp beam which is reflected\nfrom the sample surface.\n[36] The hole SDT is proportional to 1 /(δB2) for all RSA maxima with n >1. Here, δBis the full\nwidth at half maximum (FWHM) of the RSA maximum." }, { "title": "2402.04719v2.Quantum_Theory_of_Spin_Transfer_and_Spin_Pumping_in_Collinear_Antiferromagnets_and_Ferrimagnets.pdf", "content": "Quantum Theory of Spin-Transfer and Spin-Pumping in Collinear Antiferromagnets\nand Ferrimagnets\nHans Gløckner Giil and Arne Brataas\nCenter for Quantum Spintronics, Department of Physics,\nNorwegian University of Science and Technology, NO-7491 Trondheim, Norway\n(Dated: April 12, 2024)\nAntiferromagnets are promising candidates as active components in spintronic applications. They\nshare features with ferrimagnets in that opposing spin orientations exist in two or more sublattices.\nSpin transfer torque and spin pumping are essential ingredients in antiferromagnetic and ferrimag-\nnet spintronics. This paper develops an out-of-equilibrium quantum theory of the spin dynamics of\ncollinear magnets containing many spins coupled to normal metal reservoirs. At equilibrium, the\nspins are parallel or antiparallel to the easy axis. The theory, therefore, covers collinear antiferro-\nmagnets and ferrimagnets. We focus on the resulting semi-classical spin dynamics. The dissipation\nin the spin dynamics is enhanced due to spin-pumping. Spin accumulations in the normal metals\ninduce deterministic spin-transfer torques on the magnet. Additionally, each electron’s discrete spin\nangular momentum causes stochastic fluctuating torques on the antiferromagnet or ferrimagnet. We\nderive these fluctuating torques. The fluctuation-dissipation theorem holds at high temperatures,\nincluding the effects of spin-pumping. At low temperatures, we derive shot noise contributions to\nthe fluctuations.\nI. INTRODUCTION\nSpin transfer torque (STT) and spin pumping (SP)\nare essential ingredients in the generation and detection\nof spin currents and are central components in modern\nspintronics research and devices [1]. The use of mag-\nnetic insulators enables signal propagation without mov-\ning charges and could provide low-dissipation and ultra-\nfast memory devices [2]. Initially, much of spintronic\nresearch focused on the study of STT [3–5] and SP [6–\n8] in ferromagnets (FMs). Subsequently, this included\nalso works on fluctuations [9–11] and pumped magnon\ncondensates [12–14].\nUnlike FMs, whose macroscopically apparent magnetic\nproperties have been known for thousands of years, anti-\nferromagnets (AFMs) carry zero net magnetic moments\nand were elusive for some time. Even after their dis-\ncovery, AFMs were believed to have few potential ap-\nplications [15] and were disregarded in the early days\nof spintronics research. Recent theoretical and experi-\nmental findings have highlighted the potential of using\nAFMs in spintronics applications, thus starting the field\nof antiferromagnetic spintronics. Key discoveries were\nthe robustness of AFMs to external magnetic perturba-\ntions and the high resonance frequency of antiferromag-\nnetic material [16, 17]. The prediction [18] and subse-\nquent experimental detection [19] of an STT in AFMs\nsparked a massive interest in using AFMs as the active\ncomponent in spintronics devices [16, 20]. Moreover, it\nwas predicted that contrary to what was believed, anti-\nferromagnets are as efficient in pumping spin currents as\nFMs [21]. This effect was later experimentally detected in\nthe easy-axis AFM MnF 2[22]. These discoveries opened\nup the possibility of utilizing AFMs in spintronic applica-\ntions, enabling the possible fabrication of stray-field-free\ndevices operating in the THz-regime [16, 23], allowing for\nmuch faster device operation than in FMs.In recent years, the spin dynamics in AFMs have been\nexplored extensively, including the effects of disorder [24],\ngeneration of spin-Hall voltages [25], and the proper-\nties of antiferromagnetic skyrmions [26]. The spin dy-\nnamics of ferrimagnetic materials have also been stud-\nied [27]. Phenomenological models of intra- and cross-\nlattice torques were introduced in [28]. Ref. 29 fur-\nther discusses the competition between intra and cross-\nsublattice spin pumping in specific models of antiferro-\nmagnets.\nAs in antiferromagnets, ferrimagnets have opposing\nmagnetic moments. However, these moments have differ-\nent magnitudes, resulting in a net magnetization. These\nfeatures result in rich spin dynamics ranging from be-\nhavior reminiscent of antiferromagnets to ferromagnets.\nA prime example of a ferrimagnet is yttrium-iron-garnet\n(YIG). The low-energy magnon modes in YIG resemble\nmodes in ferromagnets.\nIn the study of non-equilibrium effects, the Keldysh\npath integral approach to non-equilibrium quantum field\ntheory is a powerful tool in the study of non-equilibrium\nsystems beyond linear response [30, 31]. Although most\nof the research on STT and SP utilized a semiclassical\napproach, some works have used the Keldysh framework\nin the study of spin dynamics in FMs out of equilib-\nrium [10, 11, 32–34]. Moreover, the Keldysh method was\nrecently used to formalize a fully quantum mechanical\ntheory of STT and SP, including the effects of quan-\ntum fluctuations [35]. These fluctuations have become\nincreasingly relevant with the development of new devices\noperating in the low-temperature regime. Nevertheless,\napplying the Keldysh method to derive microscopic re-\nlations for SP, STT, and fluctuating torques in an AFM\nsystem is lacking.\nIn this paper, we extend the approach of Ref. 35,\nwhich examined a ferromagnet in the macrospin approx-\nimation coupled to normal metals featuring spin andarXiv:2402.04719v2 [cond-mat.mes-hall] 11 Apr 20242\ncharge accumulation, to a similar system but instead\nfeaturing a collinear magnet with many individual spins\ncoupled at different sublattices to normal metals. Our\nstudy thus covers antiferromagnets, ferrimagnets, and\nferromagnets. We derive the spin dynamics using a\nfully quantum mechanical Keldysh non-equilibrium ap-\nproach. We find expressions for the spin transfer torque,\nspin-pumping-induced Gilbert damping, and fluctuat-\ning fields, including low-temperature shot-noise contri-\nbutions. The Gilbert damping and fluctuations contain\nboth inter-lattice and intra-lattice terms. Using Onsager\nreciprocal relations, we relate the spin pumping and spin\ntransfer coefficients. Our results enhance the knowledge\nof the microscopic expressions of STT and SP and fluctu-\nating torques in antiferromagnets and ferrimagnets cou-\npled to normal metals in the low-energy regime, where\nquantum fluctuations become essential.\nThe subsequent sections of this paper are structured\nas follows. In Sec. II, we introduce the model employed\nfor the itinerant electrons in the normal metals, the lo-\ncalized magnetic moments in the antiferromagnetic or\nferrimagnet, and the electron-magnon coupling between\nthem. We then present the key findings of this paper\nin Sec. III, including microscopic definitions of the spin\ntransfer torque, spin pumping, and fluctuating torques\nin many spin magnets, being an antiferromagnetic, fer-\nrimagnet, or ferromagnet. The derivation of an effective\nmagnon action, achieved by integrating fermionic degrees\nof freedom resulting from the interaction with normal\nmetals, is detailed in Sec. IV. The evaluation of this ef-\nfective action is then provided in Sec. V. Finally, Sec. VI\nconcludes the paper.\nII. MODEL\nWe consider a bipartite collinear magnet coupled to\nan arbitrary number of normal metal reservoirs. The\nmagnet can represent an antiferromagnet, a ferrimagnet,\nor a ferromagnet. The total Hamiltonian is\nˆH=ˆHe+ˆHem+ˆHm (1)\nin terms of the Hamiltonian describing the electrons in\nthe normal metal ˆHe, the Hamiltonian describing the in-\nteraction between the electrons and the magnet ˆHem, and\nthe Hamiltonian of the magnet ˆHm.\nThe Hamiltonian of the electrons combined with the\nHamiltonian representing the interaction between the\nelectrons and the magnet is\nˆHe+ˆHem=Z\ndrˆψ†\"\nHe+ℏ−1X\niuiσ·ˆSi#\nˆψ , (2)\nwhere ˆψ†= ( ˆψ†\n↑,ˆψ†\n↓) is the spatially dependent 2-\ncomponent itinerant electron field operator, and σis the\nvector of Pauli matrices in the 2 ×2 spin space. In the\nHamiltonian (2), ui(r) represents the spatially depen-\ndent exchange interaction between the localized spin atsiteiand the itinerant electrons. This interaction is lo-\ncalized around spin iinside the magnet. The sum over\nthe localized spins iconsists of a sum over sites in sublat-\nticeAand sublattice B, i.e.,P\ni. . .→P\na. . .+P\nb. . ..\nThe localized spin operator ˆSihas a total spin angular\nmomentum Si=ℏp\nsi(si+ 1) where siis the (unitless)\nspin quantum number of the localized spin, such that\nˆSi2=ℏ2si(si+ 1). For large si, the difference between\nSi/ℏandsiis a first-order correction, and we can ap-\nproximate Si≈ℏsi.\nThe spin-independent part of the single-particle elec-\ntron Hamiltonian is\nHe=−ℏ2\n2m∇2+Vc, (3)\nwhere Vcis the spatially dependent charge potential.\nIn the classical limit of the magnet, the spins at sublat-\nticeAare along a certain direction and the spins at sub-\nlattice Bare along the opposite direction in the ground\nstate. We will consider the semiclassical spin dynam-\nics near the instantaneous classical direction of the spins\nthat we let be along the zdirection and adiabatically\nadjust the evolution of the small deviation [10, 11, 35].\nIn the following, it is constructive to expand the in-\nteraction term to the second order in the magnet cre-\nation/annihilation operators using a Holstein-Primakoff\ntransformation,\nˆHem=ˆH0+ˆH1+ˆH2, (4)\nwhere ˆH0is the interaction with the classical magnetic\nground state and ˆH1(ˆH2) is the interaction term to the\nfirst (second) order. The classical ground state contribu-\ntion to the interaction is then\nˆH0=Z\ndrˆψ†Vsσzˆψ , (5)\nwhere the magnitude of the spatially dependent spin po-\ntential experienced by the itinerant electrons is\nVs(r) =X\nasaua(r)−X\nbsbub(r), (6)\nand oscillates rapidly with the staggered field.\nIn the macrospin approximation,P\niSican be treated\nas a giant spin in ferromagnets. Then, ui(r) becomes\nthe effective exchange interaction. Ref. 35 shows how the\nelectronic Hamiltonian ˆHecombined with the electron-\nmagnon Hamiltonian to zeroth order ˆH0become partic-\nularly transparent in ferromagnet-normal metal systems\nin terms of the scattering states of the itinerant electrons\nfor the macrospin dynamics. We generalize this approach\nto magnet-normal metal systems with individual local-\nized spins. In this picture, the electronic Hamiltonian\nremains simple, as in Ref. 35:\nˆHe+ˆH0=X\nsαϵαˆc†\nsαˆcsα, (7)3\nwhere ˆ csαannihilates an electron with spin s(s=↑or\ns=↓). The quantum number α=κnϵcaptures the\nleadκ, the transverse waveguide mode n, and the elec-\ntron energy ϵ. The electron energy consists of a trans-\nverse contribution ϵnand a longitudinal contribution\nϵ(k) =k2/2m, where kis the longitudinal momentum,\nsuch that ϵ=ϵn+ϵ(k). The eigenenergy is spin degener-\nate, since the leads are paramagnetic. Furthermore, we\nconsider identical leads such that the eigenenergy is in-\ndependent of the lead index. The system setup is shown\nL R N N AF\nFIG. 1. An antiferromagnet (AF) with conductors (N) on\neither side connected to a right (R) and a left (L) lead.\nin Fig. 1 for the case of two leads. In Eq. (7) and similar\nexpressions to follow, the sum over the scattering states\nimplies thatP\nαXsα=P\nκnR∞\nϵndϵXsκn(ϵ). In the scat-\ntering approach, the field operator is\nˆψs=X\nαˆcsαψsα, (8)\nwhere ψsα(r) is the wave function of a scattering state\nof spin sand quantum number α.\nThe Hamiltonian of the antiferromagnet is [ idenotes\na site at sublattice A(i=a) orB(i=b)]\nHm=ℏ−2X\nijJijˆSi·ˆSj−Kℏ−2X\ni\u0010\nˆSi·z\u00112\n+γµ0X\naHA\na·ˆSa+γµ0X\nbHB\nb·ˆSb, (9)\nwhere Jijis the symmetric exchange interaction, K > 0\nis the easy-axis anisotropy energy, and γ=g∗µB/ℏis\nthe (absolute value of) the effective gyromagnetic ratio,\nwhere g∗is the effective Land´ e g-factor and µBis the\nBohr magneton. In Eq. (9), HA,B\niis the external mag-\nnetic field in units of Am−1at lattice site i={a, b}, and\nµ0is the vacuum permeability, which appears because\nwe are employing SI units. In reality, HA=HBin the\npresence of a uniform external magnetic field. However,\nto illustrate and understand the physics, we allow the\nexternal fields at sublattices AandBto differ, and to\ndepend on the lattice site.\nWe consider the low-energy excitations from the semi-\nclassical ground state of the staggered spin orientation.\nTo this end, we carry out a Holstein-Primakoff expansion\nto the second order in magnon excitations at each sub-\nlattice AandBdescribed via the annihilation operators\nˆaaandˆbbas detailed in Appendix A. Introducing theraising/lowering fields as H±=Hx±iHy, the magnon\nHamiltonian becomes\nHm=E0+X\naEA\naˆa†\naˆaa+X\nbEB\nbˆb†\nbˆbb\n+ 2X\naa′Jaa′√sasa′ˆa†\naˆaa′+ 2X\nbb′Jbb′√sbsb′ˆb†\nbˆbb′\n+ 2X\nabJab√sasb[ˆaaˆbb+ ˆa†\naˆb†\nb]\n+γµ0ℏX\narsa\n2[HA\na−ˆaa+HA\na+ˆa†\na]\n+γµ0ℏX\nbrsb\n2[HB\nb−ˆb†\nb+HB\nb+ˆbb], (10)\nwhere the classical ground state energy E0is\nE0=X\naa′sasa′Jaa′+X\nbb′sbsb′Jbb′−2X\nabsasbJab\n−2KX\nis2\ni+ℏµ0X\nasaHA\naz−ℏµ0X\nbsbHB\nbz,(11)\nand is disregarded in the following.\nEA(B)\na(b)= 2X\nb(a)sb(a)Jab−X\na′(b′)sa′(b′)Ja(b)a′(b′)\n+ 2sa(b)K∓ℏγµ0HA(B)\na(b)z(12)\nis the energy of a local excitation, where the upper sign\nholds for sites on sublattice Aand the lower sign holds\nfor sites on sublattice B.\nIn the scattering basis of the electronic states, the cor-\nrections to the antiferromagnetic ground state electron-\nmagnon interaction to quadratic order in the magnet op-\nerators becomes ˆHem−ˆH0=ˆH1+ˆH2. The first-order\ncontribution of electron-magnon interaction is\nˆH1=X\naαβr\n2\nsah\nˆaaˆc†\n↓αWαβ\na↓↑ˆc↑β+ ˆa†\naˆc†\n↑αWαβ\na↑↓ˆc↓βi\n+X\nbαβr\n2\nsbh\nˆb†\nbˆc†\n↓αWαβ\nb↓↑ˆc↑β+ˆbbˆc†\n↑αWαβ\nb↑↓ˆc↓βi\n,(13)\nand describes the spin-flip scattering of the itinerant elec-\ntrons associated with creating or annihilating localized\nmagnons. The dimensionless matrix Wiis governed by\nthe exchange potential ui(r) and the scattering states\nwave functions ψsα:\nWαβ\niss‘=Z\ndrψ∗\nsα(r)siui(r)ψs‘β(r), (14)\nand is Hermitian, Wαβ\ni↑↓= [Wβα\ni↓↑]∗. The electron-magnon\ninteraction that is second order in the magnon operators4\nis\nˆH2=−X\naαβˆa†\naˆaa\nsah\nˆc†\n↑αWαβ\na↑↑ˆc↑β−ˆc†\n↓αWαβ\na↓↓ˆc↓βi\n+X\nbαβˆb†\nbˆbb\nsbh\nˆc†\n↑αWαβ\nb↑↑ˆc↑β−ˆc†\n↓αWαβ\nb↓↓ˆc↓βi\n,(15)\nwhere the matrix elements are defined in Eq. (14). We\nnote that our electron-magnon-interaction is isotropic in\nspin space and will give rise to zeroth-, first-, and second-\norder magnon terms in the Hamiltonian, i.e. ˆH0,ˆH1and\nˆH2, respectively. This is in contrast to the model used\nin Refs. 12 and 32, where only the first-order term ˆH1is\nconsidered.\nFinally, in normal metal reservoirs, the occupation of\nthe state is\n⟨c†\ns′αcsβ⟩=δαβnss′α, (16)\nwhere the 2 ×2 out-of-equilibrium distribution is\nnss′α=1\n2[fκ↑(ϵα)) +fκ↓(ϵα))]δss′\n+1\n2[fκ↑(ϵα))−fκ↓(ϵα))]uκ·σss′, (17)\nallowing for a (lead-dependent) spin accumulation in the\ndirection of the unit vector uk.f↑andf↓are general\ndistribution functions for spin-up and spin-down parti-\ncles, which generally differ for elastic or inelastic trans-\nport [35]. In equilibrium, the distribution function only\ndepends on energy,\nfeq\nκ↑(ϵ) =feq\nκ↓(ϵ) =f(ϵ−µ0), (18)\nwhere fis the equilibrium Fermi-Dirac distribution and\nµ0is the equilibrium chemical potential.\nIn inelastic transport, the spin- and charge accumula-\ntions µCandµScorrespond to chemical potential in a\n(spin-dependent) Fermi-Dirac function,\nfin\nκ↑(ϵ) =f(ϵ−µ0−µC\nκ−µS\nκ/2) (19a)\nfin\nκ↓(ϵ) =f(ϵ−µ0−µC\nκ+µS\nκ/2). (19b)\nFor notational simplicity, we define the chemical poten-\ntials\nµκ↑=µ0+µC\nκ+µS\nκ/2 (20a)\nµκ↓=µ0+µC\nκ−µS\nκ/2. (20b)\nIn the limit of small charge and spin accumulations com-\npared to the Fermi level, it can be derived that\nµC\nκ+µS\nκ\n2=Z\ndϵ\u0002\nfin\nκ↑(ϵ)−f(ϵ)\u0003\n. (21)\nIn the elastic regime, the distribution function cannot\ngenerally be described as a Fermi-Dirac function. Thedistribution function is instead given as a linear combi-\nnation of Fermi-Dirac functions in the connected reser-\nvoirs [35],\nfel\nsκ(ϵ) =X\nlRsκlf(ϵ−µl), (22)\nwhere the index lruns over the reservoirs, and Rsκlis the\nlead and spin-dependent transport coefficient for reser-\nvoirl. The transport coefficients satisfy\nX\nlRsκl= 1. (23)\nIn the elastic transport regime, it is advantageous to de-\nfine the effective charge and spin accumulations through\nµC\nκ+µS\nκ\n2=Z\ndϵ\u0002\nfel\nκ↑(ϵ)−f(ϵ)\u0003\n. (24)\nThe elastic and inelastic transport regime results in dif-\nferent results for the fluctuations in the magnetization\ndynamics of the magnet.\nHaving specified the model for the system in consider-\nation, we proceed by presenting the main results of the\npaper.\nIII. MAIN RESULTS: EQUATIONS OF MOTION\nThis section presents the main results of our work.\nOur primary result is the derivation of a Lan-\ndau–Lifshitz–Gilbert–Slonczewski (LLGS) equation for\nthe localized (normalized) spins mi=Si/Siin a gen-\neral magnet coupled to normal metal reservoirs,\n∂tmi=τb\ni+τf\ni+τsp\ni+τstt\ni (25)\nvalid for low-energy excitations when the equilibrium\nmagnetization is parallel (antiparallel) to the z-axis. The\nbulk antiferromagnet torque τb\nifor a site i={a, b}arises\nfrom contributions of anisotropy, exchange coupling, and\nexternal fields, and reads\nτb\ni=−z×\u0000\nℏ−1Eimi+γµ0Hi\u0001\n. (26)\nwhere Eiis the energy of a local excitation and Hiis\nthe applied field. Hence, the bulk torque remains unaf-\nfected by the presence of normal metal reservoirs and the\nassociated spin- and charge accumulations.\nThe spin transfer torque τstt\niis induced by spin accu-\nmulation in the normal metals, and can be expressed as\nfollows:\nτstt\ni=ℏ−1X\nκ\u0002\nβI\niκz×µS\nκ−βR\niκz×(z×µS\nκ)\u0003\n.(27)\nIn Eq. (27), the superscripts ” R,I” denote the real and\nimaginary part. The site and lead-dependent coefficients5\nβiκare expressed in terms of the microscopic scatter-\ning matrix elements defined in Eq. (14) evaluated at the\nFermi energy:\nβiκ=−2i\nsiX\nnWκnκn\ni↑↓, (28)\nand can be calculated numerically for any particular sys-\ntem configuration.\nThe spin pumping torque τsp\nicontains contributions\nfrom both sublattices and is given by\nτsp\ni=X\nj\u0002\nαR\nijz×∂tmj+αI\nijz×(z×∂tmj)\u0003\n,(29)\nwhere jruns over all sites and αijis expressed in the low-\nenergy limit using the scattering matrix elements evalu-\nated at the Fermi energy,\nαij=2π√sisjX\nκλnmWκnλm\ni↓↑Wλmκn\nj↑↓, (30)\nandαR(I)denotes the real (imaginary) part of the matrix.\nUsing the Onsager reciprocal relations in Appendix B, we\nfind that the spin transfer torque and spin pumping are\nrelated in the case of the most relevant case of a single\nreservoir,\nX\njαij=βi. (31)\nFinally, the fluctuating torque τf\niis expressed in terms\nof a fluctuating transverse field Hf\ni,\nτf\ni=−γµ0z×Hf\ni. (32)\nThe fluctuating field exhibit interlattice and intralattice\ncorrelators ⟨HµiHνj⟩, where µ, ν={x, y}:\n2√sisjγ2µ2\n0⟨Hf\nxiHf\nxj⟩= ImΣK\nij+ 4Im ˜Σ↑↓ij (33a)\n2√sisjγ2µ2\n0⟨Hf\nxiHf\nyj⟩=−ReΣK\nij−4Re˜Σ↑↓ij (33b)\n2√sisjγ2µ2\n0⟨Hf\nyiHf\nyj⟩= ImΣK\nij−4Im˜Σ↑↓ij, (33c)\nwhere the time arguments tandt′of the fields and the\nrelative time argument ( t−t′) of the self energies are\nomitted for simplicity. The self-energy Σ is due to charge\nand longitudinal spin accumulations in the normal metals\nand is nonzero even in equilibrium. It is conveniently\nwritten as a product of a frequency-dependent quantity\nπ(ω) and a site and scattering states dependent quantity\nσij[35]:\nΣK\nij(ω) =i\nℏX\nκλσijκλπκλ(ω), (34)\nwith\nπκλ(ω) =−2Z\ndϵ[2n↑↑κ(ϵ)n↓↓λ(ϵ+ℏω)\n−n↑↑κ(ϵ)−n↓↓λ(ϵ+ℏω)] (35a)\nσijκλ=2π√sisjX\nnmWκnλm\ni↓↑Wλmκn\nj↑↓. (35b)Conversely, the self-energy matrices ˜Σ↑↓are due to trans-\nverse spin accumulation in the normal metals and, as a\nresult, vanish in equilibrium. Analogous to the decom-\nposition in Eq. (34), we write\n˜ΣK\n↑↓ij=−i\nℏX\nκλ˜σ↑↓ijκλ˜πκλ(ω), (36)\nwhere\n˜π↑↓(ω) =−4Z\ndϵn↑↓κ(ϵ)n↑↓λ(ϵ+ℏω) (37a)\n˜σ↑↓ijκλ=−π√sisjX\nnmWκnλm\ni↓↑Wλmκn\nj↓↑, (37b)\nThe noise matrices π(ω) and ˜ π(ω) are similar to what\nwas found in Ref. 35, and are calculated in the equi-\nlibrium, elastic, and inelastic transport regime in Sec.\nV B. Crucially, the shot noise differs on various sites, due\nto the site-dependence of σand ˜σ. At equilibrium, the\nfluctuation-dissipation theorem holds, e.g.\n2siγ2µ2\n0⟨Hf\nµiHf\nνi⟩=δµναii4kBTξ\u0012ℏω\n2kBT\u0013\n, (38)\nwhere ξ(x) =xcothx.\nIn the next section, we discuss the Keldysh action of\nthe model presented in this section and derive an effec-\ntive action by integrating out the fermionic degrees of\nfreedom.\nIV. KELDYSH THEORY AND EFFECTIVE\nACTION\nIn this section, we derive the semiclassical spin dynam-\nics by using an out-of-equilibrium path integral formal-\nism [30]. We introduce the closed contour action Sand\nthe partition function Z,\nZ=Z\nD[¯aa¯bb¯c↑c↑¯c↓c↓]eiS/ℏ. (39)\nThe action Sconsists of contributions from the localized\nmagnetic excitations aandb, and the spin-up c↑and\nspin-down electrons c↓from the scattering states. We\nwill integrate out the fermion operators and get an effec-\ntive action for the magnetic excitations aandb, which\nincludes effective transverse and longitudinal fields that\narise from the charge and spin accumulations in the nor-\nmal metals.\nWe follow Ref. 30 and replace the fields in Eq. (39)\nwith ” ±” fields residing on the forward and backward\npart of the Schwinger-Keldysh contour. The action in\nthe±basis is given in Appendix C. These fields are not\nindependent of each other and can be Keldysh rotated\ninto a new basis that takes into account the coupling\nbetween them. The rotated fields have the advantage\nof suggesting a transparent physical interpretation, cor-\nresponding to the semiclassical equations and quantum\ncorrections.6\nA. Keldysh action\nFor magnons, the classical ( cl) and quantum ( q) fields\nare defined linear combinations of the ±-fields, as de-\nscribed in detail in Appendix C. In Keldysh space, it is\nconvenient to also introduce the matrices\nγq=\u0012\n0 1\n1 0\u0013\nγcl=\u0012\n1 0\n0 1\u0013\n. (40)\nThe Keldysh rotated magnon action becomes\nSm=X\nat¯aq\na(iℏ∂t−EA\na)acl\na+X\nbt¯bq\nb(iℏ∂t−EB\nb)bcl\nb\n+X\nataq\na(iℏ∂t−EA\na)¯acl\na+X\nbtbq\nb(iℏ∂t−EB\nb)¯bcl\nb\n−2X\naa′t√sasa′Jaa′\u0002\n¯aq\naacl\na′+h.c.\u0003\n−2X\nbb′t√sbsb′Jbb′\u0002¯bq\nbbcl\nb′+h.c\u0003\n−2X\nabt√sasbJab\u0002\n¯aq\na¯bcl\nb+ ¯acl\na¯bq\nb+h.c.\u0003\n−γµ0ℏX\nat√sa\u0002\nHA\na−aq\na+HA\na+¯aq\na\u0003\n−γµ0ℏX\nbt√sb\u0002\nHB\nb−¯bq\nb+HB\nb+bq\nb\u0003\n, (41)where we wrote the time integral as a sum for compact\nnotation. In Eq. (41), h.c.denotes the hermitian conju-\ngate of the previous term. The fermion action becomes\nSe+S0=X\nst¯Csγcl(iℏ∂t−ϵ)Cs, (42)\nwhere we introduced vector notation for the 1 /2-fields,\n¯Cs= (¯c1\nsα,¯c2\nsα), and ϵis a diagonal matrix containing\nthe single-particle energies of the electrons. The Keldysh\nrotated first-order electron-magnon interaction is\nS1=−X\nat\nαβ1√sah\nacl\naWαβ\na↓↑¯C↓αγclC↑β+aq\naWαβ\na↓↑¯C↓αγqC↑β+h.c.i\n−X\nbt\nαβ1√sbh\n¯bcl\naWαβ\nb↓↑¯C↓αγclC↑β+¯bq\nbWαβ\nb↓↑¯C↓αγqC↑β+h.c.i\n, (43)\nand, finally, the second-order term reads\nS2=X\nat\nαβ1\n2sah\nWαβ\na↑↑\u0000¯AaγclAa¯C↑αγclC↑β+¯AaγqAa¯C↑αγqC↑β\u0001\n−Wαβ\na↓↓\u0000¯AaγclAa¯C↓αγclC↓β+¯AaγqAa¯C↓αγqC↓β\u0001i\n−X\nbt\nαβ1\n2sbh\nWαβ\nb↑↑\u0000¯BbγclBb¯C↑αγclC↑β+¯BbγqBb¯C↑αγqC↑β\u0001\n−Wαβ\nb↓↓\u0000¯BbγclBb¯C↓αγclC↓β+¯BbγqBb¯C↓αγqC↓β\u0001i\n,\n(44)\nwhere the magnon q/cloperators are consolidated in vec-\ntors ¯Aaand ¯Bb. The Keldysh rotated action proves to\nbe well-suited for the computation of an effective magnon\naction, a topic we delve into in the following section.B. Integrating out the fermionic degrees of\nfreedom\nFor the itinerant electrons in the normal metal and\nantiferromagnet, the total effective electron action is7\nSe,tot=Se+Sem, and can be expressed as\nSe,tot=X\nss′tt′¯Cs,tG−1\nss′,tt′Cs′,t′, (45)\nwhere the interacting Green function Gis given in terms\nof the noninteracting Green function G0and interaction\nterms as\nG−1=G−1\n0+˜W1+˜W2. (46)\nHere, ˜W1contains the first-order magnon operators on\nboth sublattices,\n˜W1=δ(t−t′)\u0002\nWA\n1+WB\n1\u0003\n, (47)\nwhere WA\n1andWB\n1are spin flip operators:\nWA\n1=−X\nxa1√saγx\u0002\nWa↑↓¯ax\naσ++Wa↓↑ax\naσ−\u0003\n(48a)\nWB\n1=−X\nxb1√sbγx\u0002\nWb↑↓bx\nbσ++Wb↓↑¯bx\nbσ−\u0003\n.(48b)\nIn Eq. (48), the variable x={cl, q}represents a Keldysh\nspace index, and σ±are the usual raising and lowering\nPauli matrices. Similarly, ˜W2contains the magnon oper-\nators to quadratic order for both sublattices,\n˜W2=δ(t−t′)\u0002\nWA\n2−WB\n2\u0003\n, (49)\nwith WA\n2andWB\n2given by\nWA\n2=X\naxy1\n2sa¯ax\naγxay\naγy\u0012\nWa↑↑ 0\n0−Wa↓↓\u0013\n(50a)\nWB\n2=X\nbxy1\n2sb¯bx\nbγxby\nbγy\u0012\nWb↑↑ 0\n0−Wb↓↓\u0013\n, (50b)\nwhere the spin structure is explicitly written out as a ma-\ntrix. The matrices WA(B)\n1 andWA(B)\n2 have a structure\nin the scattering states space from Wa(b), spin space from\nthe Pauli matrices, and Keldysh space from γx. The in-\nverse free electron Green function G−1\n0from Eq. (46) has\nthe conventional causality structure in Keldysh space,\nwith a retarded ( R), advanced ( A), and Keldysh ( K)\ncomponent:\nG−1\n0=\u0012\n[GR\n0]−1[GK\n0]−1\n0 [ GA\n0]−1\u0013\n, (51)\nand has equilibrium components that are diagonal in\nboth spin space and in the scattering states space,\n[G−1\n0]R(A)\nαβ,ss′=δαβδss′δ(t−t′)(iℏ∂t−ϵα±iδ),(52)\nwhere the upper sign corresponds to the retarded compo-\nnent, while the lower sign is applicable to the advanced\ncomponent. The Keldysh component includes informa-\ntion about the distribution function, and will be dis-\ncussed below, when we Fourier transform the Green func-\ntions.From the effective electron action in Eq. (45), it is\nevident that the partition function of Se,tottakes on a\nGaussian form with respect to the fermionic operators.\nHence, the fermionic integral in the partition function\ncan be evaluated exactly, with an inconsequential pro-\nportionality constant being disregarded:\nZ\nD[C]eiSe,tot/ℏ= eTr[ln[1+G0˜W1+G0˜W2]]. (53)\nIn Eq. (53), we have used the short-hand notation for\nthe functional integral measure of all fermionic states,\nD[C] =D[¯C↑C↑¯C↓C↓]. We have absorbed a normaliza-\ntion constant into the functional integral measure for sim-\nplicity. We note that as a consequence of the continuity\nof the time coordinate and scattering states energy that\nwe are employing, the unit matrix is a delta function in\ntime and energy, 1 ≡δ(t−t′)δ(ϵα−ϵβ), and thus quanti-\nties inside the logarithm carries dimension J−1s−1. The\ntrace, on the other hand, is an integral operator with unit\nJ s. As long as one interprets the logarithm in terms of\nits Taylor expansion, this does not lead to any problems,\nas the exponent of Eq. (53) becomes dimensionless for all\nterms in the expansion. The exponent is interpreted as\nan additional contribution to the magnon action,\ni\nℏSeff= Trh\nln[1 + G0˜W1+G0˜W2]i\n. (54)\nThe way forward is to treat this interaction as a pertur-\nbation, expanding the logarithm in first and second-order\ncontributions and disregarding higher-order terms,\nSeff≈ −iℏTrh\nG0˜W1i\n−iℏTrh\nG0˜W2i\n+iℏ\n2Trh\nG0˜W1G0˜W1i\n. (55)\nTo evaluate the trace in these terms, it is convenient to\nFourier transform all quantities from the time domain\nto the energy domain. This diagonalizes the noninter-\nacting Green functions, making calculations much more\nstraightforward.\nC. Fourier representation\nThe paper employs the Fourier transform convention\ndefined in Appendix D. In Fourier space, the fermion\nequilibrium Green function components are particularly\nsimple:\n[G0]R(A)\nαβ,ss′(ω) =δαβδss′(ℏω−ϵ±iδ)−1, (56)\nwhere δ > 0 is an infinitesimal quantity ensuring con-\nvergence. The Keldysh component accounts for non-\nequilibrium phenomena though the spin-dependent dis-\ntribution nss′αdefined in Eq. (17),\n[G0]K\nαβ,ss′(ω) =−2πiδαβδ(ℏω−ϵα) [δss′−2nss′α].(57)8\nThe Keldysh component has off-diagonal terms in spin\nspace if the distribution function nss′αhas off-diagonal\nelements, i.e. if there is a transverse spin accumulation\nin the normal metals.\nV. NON-EQUILIBRIUM SPIN DYNAMICS\nHaving derived the effective action as expressed in Eq.\n(55), we proceed by evaluating the traces and delving\ninto the resultant terms. The discussion unveils effective\nlongitudinal and transverse fields, which we ascribe to\nspin transfer torque and spin pumping originating from\nthe normal metal reservoirs.\nA. First-order contribution\nEvaluating the trace in the first order term in Eq.\n(55) corresponds to summing over the diagonal elements\nin spin space and Keldysh space, integrating over both\ntime variables, and summing over the space of scattering\nstates, we find\n−iℏTrh\nG0˜W1i\n=−X\naα2√saWαα\na↑↓n↓↑αZ\ndt¯aq\na(t)\n−X\naα2√saWαα\na↓↑n↑↓αZ\ndtaq\na(t)\n−X\nbα2√saWαα\nb↑↓n↓↑αZ\ndtbq\nb(t)\n−X\nbα2√saWαα\nb↓↑n↑↓αZ\ndt¯bq\nb(t).(58)\nHere, we have used the general Green function identity\nGR(t, t)+GA(t, t) = 0 [30], and written the time integra-\ntion explicitly. Comparing the first-order contribution in\nEq. (58) with the magnon action in Eq. (41), we observe\nthat the first-order effect of the spin accumulation in the\nnormal metal is equivalent to an effective deterministic\ntransverse magnetic field Hstt\ni, which act on a localized\nspin at site i={a, b}in the antiferromagnet. The ”stt”\nsuperscript indicates that this field will take the form of\na spin transfer torque, which will be elaborated on be-\nlow. The magnitudes of these effective transverse fields\nare given by\nγµ0Hstt\ni−=2\nsiℏX\nαWαα\ni↓↑n↑↓α (59a)\nγµ0Hstt\ni+=2\nsiℏX\nαWαα\ni↑↓n↓↑α, (59b)\nwhich implies that the Cartesian components read\nγµ0Hstt\nix=2\nsiℏX\nαRe\u0002\nWαα\ni↑↓n↓↑α\u0003\n(60a)\nγµ0Hstt\niy=2\nsiℏX\nαIm\u0002\nWαα\ni↑↓n↓↑α\u0003\n. (60b)Recalling that the spin accumulation is given by Eq. (24)\nand Eq. (21), we write the effective fields from Eq. (60)\nin the conventional spin transfer torque form:\nγµ0Hstt\ni=1\nℏX\nκ\u0002\nβR\niκz×µS\nκ+βI\niκz×(z×µS\nκ)\u0003\n,(61)\nwhere the appearance of zis a consequence of our the-\nory being restricted to small deviations for the equilib-\nrium magnetization ±z. This results in the spin transfer\ntorque given in Eq. (27). In Eq. (61), the superscripts\n”R” and ” I” denote the real and imaginary parts and the\nlead- and site-dependent constants βiκhave been intro-\nduced as sums over the transverse modes of the scattering\nmatrix elements,\nβiκ=−2i\nsiX\nnWκnκn\ni↑↓, (62)\nand where we have assumed that the transverse spin dis-\ntribution functions n↑↓andn↓↑are only significant close\nto the Fermi surface, such that the scattering states ma-\ntrix elements are well approximated by their value at\nthe Fermi surface. The expression for the spin transfer\nfield in Eq. (61) is valid in both the elastic and inelastic\nregime, and vanishes in equilibrium. We note that the\ncoefficient βiκ, for i={a, b}, depends not only on the\npotential at lattice site ibut also indirectly of all lattice\nsites on both sublattices through the scattering states.\nTo the lowest order, the sublattice magnetizations are\nparallel and antiparallel to the z-axis, mA≈zand\nmB≈ −z. Thus, to the lowest order in the magnon\noperators, the expressions for the transverse fields are\nambiguous, and we can write the transverse field in Eq.\n(61) in terms of mAormB. To the lowest order in the\nmagnon operators, the Keldysh technique cannot be used\nto identify which sublattice the transverse fields in Eq.\n(58) originate from.\nB. Second order contribution\nThe second order contribution in Eq. (55) has contri-\nbutions from ˜W2,\nS21=−iℏTrh\nG0˜W2i\n, (63)\nas well as a contribution from ˜W1,\nS22=iℏ\n2Trh\nG0˜W1G0˜W1i\n. (64)\nProceeding in a manner analogous to the treatment of\nthe first-order term, the trace in S21is evaluated:9\nS21=−X\naαπ\nsa\u0002\nWαα\na↑↑(1−2n↑↑α)−Wαα\na↓↓(1−2n↓↓α)\u0003Z\ndt¯Aa(t)γqAa(t)\n+X\nbαπ\nsb\u0002\nWαα\nb↑↑(1−2n↑↑α)−Wαα\nb↓↓(1−2n↓↓α)\u0003Z\ndt¯Bb(t)γqBb(t). (65)\nFrom Eq. (41), it is apparent that the second-order terms in S21are equivalent with a longitudinal magnetic field,\nwith magnitude\nγµ0HA21\niz=−π\nℏsiX\nα\u0002\nWαα\ni↑↑(1−2n↑↑α)−Wαα\ni↓↓(1−2n↓↓α)\u0003\n, (66)\nwhich, in this reference frame, renormalizes the energies of localized magnon excitations. However, such longitudinal\nmagnetic fields should not affect the spin dynamics since they, in the instantaneous reference field, correspond to\ncontributions to the total free energy proportional to S2\ni.\nThe final contribution S22to the effective action contains inter-lattice and intra-lattice terms and can be written\ncompactly by introducing a field di={aa,¯bb}and summing over the two field components, i.e.P\nidi=P\naaa+P\nb¯bb:\nS22=Z\ndtdt′X\nxx′ijiℏ\n2√sisjTrh\nG0,↑↓(t′, t)Wi↓↑γxG0,↑↓(t, t′)γx′Wj↓↑i\ndx\ni(t)dx′\nj(t′)\n+Z\ndtdt′X\nxx′ijiℏ\n2√sisjTrh\nG0,↓↑(t′, t)Wi↑↓γxG0,↓↑(t, t′)γx′Wj↑↓i\n¯dx\ni(t)¯dx′\nj(t′)\n+Z\ndtdt′X\nxx′ijiℏ\n2√sisjTrh\nG0,↓↓(t′, t)Wi↓↑γxG0,↑↑(t, t′)γx′Wj↑↓i\ndx\ni(t)¯dx′\nj(t′)\n+Z\ndtdt′X\nxx′ijiℏ\n2√sisjTrh\nG0,↑↑(t′, t)Wi↑↓γxG0,↓↓(t, t′)γx′Wj↓↑i\n¯dx\ni(t)dx′\nj(t′), (67)\nwhere the trace is taken only over the 2 ×2 Keldysh space and the space of scattering states α. The interlattice terms,\ni.e.d=d′, are discussed in Ref. 35 for a macrospin ferromagnet. Here, we summarize this discussion and highlight\nthe addition of the inter-lattice terms not present in the macrospin ferromagnet.\nEvaluating the trace in the first and second line of Eq. (67), we note that only the Keldysh component has off-\ndiagonal elements in spin space, and find a contribution only from x=x′=q,\n˜Sqq\n22=ℏZ\ndtdt′X\nijh\ndq\ni(t)˜ΣK\n↑↓ij(t, t′)dq\nj(t′)i\n(68a)\n˜S¯q¯q\n22=ℏZ\ndtdt′X\nijh\n¯dq\ni(t)˜ΣK\n↓↑ij(t, t′)¯dq\nj(t′)i\n, (68b)\nwhere the self-energies are\n˜ΣK\n↑↓ij(t−t′) =−2i\nℏ2√sisjX\nαβn↑↓αn↑↓βWαβ\ni↓↑Wβα\nj↓↑ei(ϵα−ϵβ)(t−t′)/ℏ(69a)\n˜ΣK\n↓↑ij(t−t′) =−2i\nℏ2√sisjX\nαβn↓↑αn↓↑βWαβ\ni↑↓Wβα\nj↑↓ei(ϵα−ϵβ)(t−t′)/ℏ. (69b)\nThe reasoning behind identifying this self-energy as a Keldysh component is that it couples the quantum components\nof the fields, see Eq. (68). The terms in Eq. (68) do not have a direct analog in the magnon action in Eq. (41), and\ninterpreting these will be the subject of Sec. V C. The self-energies in Eq. (69) are invariant under a joint time and\nlattice site reversal, i.e. ˜Σij(t−t′) =˜Σji(t′−t). Moreover, due to the properties n↑↓=n∗\n↓↑andWαβ\ni↑↓= [Wβα\ni↓↑]∗, we\nsee that the self-energies are related by ˜ΣK\n↑↓ij(t−t′) =−[˜ΣK\n↓↑ij(t−t′)]∗, which will be important later.\nDisregarding terms of the order kBT/ϵFandµs/ϵF[35], we find that the Fourier-transformed self-energy becomes\n˜ΣK\n↑↓ij(ω) =−i\nℏX\nκλ˜σ↑↓ijκλ˜πκλ(ω), (70)10\nwhere\n˜πκλ(ω) =−4Z\ndϵn↑↓κ(ϵ)n↑↓λ(ϵ+ω) ˜ σ↑↓ijκλ=−π√sisjX\nnmWκnλm\ni↓↑Wλmκn\nj↓↑, (71)\nand where the matrix elements Ware evaluated at the Fermi surface. This is a straightforward generalization of\nthe macrospin ferromagnet case, with the addition of shot-noise contributions from inter-lattice and intra-lattice\ninteractions between different lattice sites. We can evaluate the quantity ˜ πκλ↑↓(ω) by using Eq. (17):\n˜π↑↓κλ(ω) =−uκ−uλ−Z\ndϵ[f↑κ(ϵ)−f↓κ(ϵ)] [f↑λ(ϵ+ℏω)−f↓λ(ϵ+ℏω)] (72)\nwhere we introduced the conventional ”lowering” vector u−=ux−iuy. This can be computed in equilibrium, elastic,\nand inelastic scattering cases, and results exactly similar to those in Ref. 35.\nWe now turn our attention to the third and fourth lines of the second-order action in Eq. (67). The contributions\nfrom the two lines are equal, which is evident from interchanging summation indices and rearranging terms. Their\ntotal contribution to the action S22can be split into contributions S¯qq\n22,S¯qcl\n22, and S¯clq\n22. The contribution Sclclvanishes,\ndue to the quantity GR(t′−t)GR(t−t′) being nonzero only for t=t′, which has measure zero, and similarly for\nGA. This ensures that the action satisfies the general requirement S[ϕcl, ϕq= 0] = 0 [30]. Introducing, for notational\nconvenience, the vector ¯Di=\u0000¯dcl¯dq\u0001\n, we find\nS¯qq\n22+S¯qcl\n22+S¯clq\n22=ℏZ\ndtdt′X\nij¯Di(t)ˆΣij(t−t′)Dj(t′), (73)\nwhere the self-energy matrix has structure in Keldysh space and in the sublattice space,\nˆΣij(t−t′) =\u0012\n0 ΣA(t−t′)\nΣR(t−t′) ΣK(t−t′)\u0013\nij, (74)\nand its components are given by\nΣK\nij(t−t′) =2i√sisjℏ2X\nαβ(n↑↑α+n↓↓β−2n↑↑αn↓↓β)Wαβ\ni↓↑Wβα\nj↑↓ei(ϵα−ϵβ)(t−t′)/ℏ(75a)\nΣR\nij(t−t′) =2i√sisjℏ2θ(t−t′)X\nαβ(n↑↑α−n↓↓β)Wαβ\ni↓↑Wβα\nj↑↓ei(ϵα−ϵβ)(t−t′)/ℏ(75b)\nΣA\nij(t′−t) =−2i√sisjℏ2θ(t−t′)X\nαβ(n↑↑α−n↓↓β)Wαβ\ni↓↑Wβα\nj↑↓ei(ϵα−ϵβ)(t−t′)/ℏ. (75c)\nThe Keldysh component of this self-energy has the symmetry [ΣK\nij(t−t′)]∗=−ΣK\nji(t′−t). Imperatively, as a\nconsequence of this symmetry, the quantities\nΣK\nij(t−t′)−ΣK\nji(t′−t) = 2Re\u0002\nΣK\nij(t−t′)\u0003\n(76)\niΣK\nij(t−t′) + iΣK\nji(t′−t) =−2Im\u0002\nΣK\nij(t−t′)\u0003\n, (77)\nare real numbers, which will be important in the next section. We proceed by a similar analysis to what was done\nwith ˜Σ, writing it in terms of a shot-noise matrix. We assume that the matrices Wcan be approximated by their\nvalue on the Fermi surface, and write\nΣK\nij(ω) =i\nℏX\nκλσijκλπκλ(ω), (78)\nwhere we introduced the matrices\nπκλ(ω) =−2Z\ndϵ[2n↑↑κ(ϵ)n↓↓λ(ϵ+ℏω)−n↑↑κ(ϵ)−n↓↓λ(ϵ+ℏω)] σijκλ=2π√sisjX\nnmWκnλm\ni↓↑Wλmκn\nj↑↓.(79)\nThe matrix π(ω) can also be evaluated in equilibrium, and for elastic and inelastic scattering, and the results are\nagain exactly similar to those in Ref. 35.\nwhere ξ(x) =xcothxis an asymptotically linear function for high xandpss′κ= (1−uzκ)/2+uzκδssis a projection11\nfactor introduced for notational convenience.\nComparing with the magnetic action in Eq. (41), we\nnotice that the terms with the retarded and advanced\nself-energies are equivalent with longitudinal fields, which\nwe in the following will show consists of dissipative\nGilbert-like terms and non-dissipative field-like terms.\nFourier transforming and applying the identity (D5), the\nretarded and advanced self-energies from Eq. (75b) and\nEq. (75c) become\nΣR,A\nij=−2√sisjℏX\nαβn↑↑α−n↓↓β\nℏω+ϵα−ϵβ±iδWαβ\ni↓↑Wβα\nj↑↓.(80)\nThis self-energy has equilibrium contributions as well\nas non-equilibrium contributions, however, the non-\nequilibrium contributions scale as µ↑↑/ϵFandµ↓↓/ϵFand\nare disregarded in the following. The equilibrium part of\nEq. (80) becomes particularly transparent when expand-\ning to first order in the frequency ω:\nΣR/A\n↑↓ij(ω)≈ΣR/A\n↑↓ij(ω= 0)±iωαij, (81)\nwhere we introduced the frequency-independent matrix\nelement\nαij=2π√sisjX\nαβ[−f′(ϵα)]δ(ϵα−ϵβ)Wαβ\ni↓↑Wβα\nj↑↓,(82)\nwhich can be approximated to\nαij=2π√sisjX\nκλnmWκnλm\ni↓↑Wλmκn\nj↑↓, (83)\nwhere the scattering states matrix elements are evalu-\nated at the Fermi surface. We note from the identity\n[Wαβ\ni↑↓]∗=Wβα\ni↓↑thatαis a Hermitian matrix in the space\nof lattice sites, i.e. [ αij]∗=αji. The zeroth order term\nin frequency is\n[S¯qcl\n22+S¯clq\n22]0=ℏX\nijZ\ndω¯dq\ni(ω)ΣR\n↑↓ij(0)dcl\nj(ω)\n+ℏX\nijZ\ndω¯dcl\ni(ω)ΣA\n↑↓ij(0)dq\nj(ω),(84)which is a constant longitudinal field that plays no role\nin the instantaneous reference frame, as discussed above.\nThe first-order term in frequency is finite even in equi-\nlibrium,\n[S¯qcl\n22+S¯clq\n22]1=ℏX\nijαijZ\ndt¯Diγq∂tDj, (85)\nand takes the form of a Gilbert damping term, includ-\ning both inter-lattice and intra-lattice contributions. The\nspin transfer torque coefficient αand the spin pumping\ncoefficient βare related to each other as a consequence of\nthe Onsager reciprocal relations [36]. In Appendix B we\nderive this relation, which is given in Eq. (B11), and de-\nrive an optical theorem relating the scattering matrices,\ngiven in Eq. (B12).\nSummarizing this section, we have found that the cor-\nrections to the magnon action Smin the presence of spin\nand charge accumulations in surrounding normal metals\nisS1+S21+S¯qcl\n22+S¯clq\n22+˜Sqq\n22+S¯qq\n22, and found that the\nfirst three of these contributions appear like magnetic\nfields and (in the low-frequency limit) like Gilbert-like\ndamping terms in the effective magnon action. Impor-\ntantly, we find both longitudinal and transverse fields in\nthe general case. The last two contributions to the ac-\ntion consist of coupled quantum fields and are the result\nof purely quantum effects. These terms are the subject\nof the next section.\nC. Fluctuating fields\nFrom the effective action in the last section, we were\nable to associate the ( q, cl) and ( cl, q) terms with longi-\ntudinal fields by comparing them with the magnon ac-\ntion in Eq. (41). Now, we must address the issue of\nhow to interpret the ( q, q) terms, which lack an ana-\nlog in the action described in Eq. (41). In this section,\nwe derive fluctuating forces from these terms by employ-\ning a Hubbard-Stratonovich (HS) transformation on the\nquadratic fields in the effective action, introducing aux-\niliary fields in the process. Commencing with the con-\ntribution from the term S¯qq\n22, we introduce the complex\nauxiliary field h¯qq\ni(in units of inverse second) via a con-\nventional Hubbard–Stratonovich transformation:\neiS¯qq\n22/ℏ= exp\u0014Z\ndtdt′X\nij¯dq\ni(t)iΣK\nij(t−t′)dq\nj(t′)\u0015\n=1\ndet [−iΣK]ZY\niD[h¯qq\ni] exp\u0014\niZ\ndtX\nih¯qq\ni(t)¯dq\ni(t) +h.c.−Z\ndtdt′X\nij¯h¯qq\ni(t)[−iΣK\nij(t−t′)]−1hqq\nj(t′)\u0015\n,\n(86)\nwhere a shorthand notation for the measure was in-\ntroduced as D[h¯qq\ni] = Π k\b\nd[Imh¯qq\ni(tk)]d[Reh¯qq\ni(tk)]/π\t\n,where kis the index used to order the discretization of12\nthe time coordinate. From the Gaussian form of Eq. (86),\nthe correlators of the auxiliary field can be identified as\n⟨h¯qq\ni(t)⟩= 0 (87a)\n⟨h¯qq\ni(t)h¯qq\nj(t′)⟩= 0 (87b)\n⟨¯h¯qq\ni(t)h¯qq\nj(t′)⟩=−iΣK\nij(t−t′). (87c)\nThe second term in the exponent is quadratic in the newfields, and gives no contribution to the magnon action,\nwhile the first term is linear in the magnon field diand is\ninterpreted as an effective transverse field in the magnon\naction.\nThe contribution from the terms ˜Sqq\n22+˜S¯q¯q\n22is HS trans-\nformed by performing an unconventional transformation\nin the two complex fields ˜hqq\niand˜hqq\niseparately:\nei˜S22/ℏ=1q\ndet [−2i˜ΣK\n↓↑]q\ndet [−2i˜ΣK\n↑↓]ZY\niD[˜hqq\ni] exp\"\niZ\ndtX\ni\u0010\n˜hqq(t)˜hqq(t)\u0011\ni\u0012\nd(t)\n¯d(t)\u0013\ni+h.c.\n−Z\ndtdt′X\nij\u0010\n˜hqq(t)˜hqq(t)\u0011\ni \n0 −i˜ΣK\n↓↑(t−t′)\n−i˜ΣK\n↑↓(t−t′) 0!−1\nij ˜hqq(t′)\n˜hqq(t′)!\nj#\n,(88)\nagain interpreting the exponent as an effective action in-\ncluding the field ˜hqq\ni, which has the correlators\n⟨˜hqq\ni(t)⟩= 0 (89a)\n⟨˜hqq\ni(t)˜hqq\nj(t′)⟩=−i˜ΣK\n↑↓ij(t−t′) (89b)\n⟨˜hqq\ni(t)˜hqq\nj(t′)⟩=−i˜ΣK\n↓↑ij(t−t′) (89c)\n⟨˜hqq\ni(t)˜hqq\nj(t′)⟩= 0. (89d)\nWe remark that the unconventional form of the Hubbard-\nStratonovich decoupling leads to non-zero correlators for\nequal fields, as opposed to the conventional approach\nwhere the non-zero correlators involve one field being\nthe complex conjugate of the other. The fields h¯qqand\n˜hqqare interpreted as fluctuating transverse fields with,\nin general, different amplitudes depending on the lattice\nsite, but with correlators between lattice sites. Compar-\ning the effective action in Eq. (86) and Eq. (88) with the\nmagnon action in Eq. (41), the components of the total\nfluctuating field Hfcan be identified as\nγµ0Hf\n+,i=−1√sih\n2˜hqq\ni+h¯qq\nii\n(90a)\nγµ0Hf\n−,i=−1√sih\n2˜hqq\ni+¯h¯qq\nii\n. (90b)In this expression, the factor of 2 arises from the uncon-\nventional nature of the Hubbard-Stratonovich transfor-\nmation in Eq. (88). The correlators between the Carte-\nsian components of the fluctuating field can be calculated\nusing Eq. (89) and Eq. (90),\n2√sisjγ2µ2\n0⟨Hf\nxiHf\nxj⟩= ImΣK\nij+ 4Im ˜Σ↑↓ij (91a)\n2√sisjγ2µ2\n0⟨Hf\nxiHf\nyj⟩=−ReΣK\nij−4Re˜Σ↑↓ij (91b)\n2√sisjγ2µ2\n0⟨Hf\nyiHf\nyj⟩= ImΣK\nij−4Im˜Σ↑↓ij, (91c)\nfrom which we conclude that the correlators in the fluc-\ntuating field Hfare real numbers. In Eq. (91), we omit-\nted the time arguments for notational simplicity. Fur-\nthermore, it is evident that for i=jandt=t′, the\ncorrelators in Eq. (91a) and (91c) are positive, aligning\nwith the conditions expected for representing the vari-\nance of a real field.13\nD. Equations of motion\nAfter HS decoupling the qqcomponents, the effective action reads\nSeff=−γℏµ0Z\ndt\"X\na√sa\u0000\nHstt\na++Hf\na+\u0001\n¯aq\na(t) +X\nb√sb\u0000\nHstt\nb−+Hf\nb−\u0001¯bq\nb(t) +h.c.#\n+ℏZ\ndt\"X\naa′βaa′¯aq\na∂tacl\na+X\nabβabaq\na∂tbcl\nb+X\nbaβba¯bq\nb∂t¯acl\na+X\nbb′βbb′¯bq\nb∂tbcl\nb′+h.c.#\n. (92)\nHaving cast the total action Sm+Seffin a form that is linear in the quantum fields ¯ aqand¯bqand their complex\nconjugates, we can integrate over these fields in the partition function, producing the functional delta function imposing\nthe semiclassical equations of motion for the fields aclandbcl[30]. Using acl\na=S+a/(ℏ√sa) and ¯bcl\nb=S+b/(ℏ√sb) in\nthe semiclassical limit, we find the coupled equations of motion:\ni∂tSi+=ℏ−1EiSi++siℏµ0γ\u0000\nHi++Hf\ni++Hstt\ni+\u0001\n−X\njβij∂tSj+, (93)\nas well as its complex conjugated counterpart. Both in the definition of this field and in Eq. (93), the upper sign holds\nfor sublattice A, and the lower sign holds for sublattice B. We find the Cartesian components by taking the real and\nimaginary parts and divide with ℏsito find an equation for the vector mi=Si/(ℏsi),\n∂tmi=τb\ni+τf\ni+τsp\ni+τstt\ni, (94)\nwhere\nτb\ni=−z×\u0000\nℏ−1Eimi+γµ0Hi\u0001\n(95a)\nτf\ni=−γµ0z×Hf\ni (95b)\nτstt\ni=−γµ0z×Hstt\ni (95c)\nτsp\ni=X\njReβijz×∂tmj+X\njImβijz×(z×∂tmj) (95d)\nis microscopic expressions for the bulk torque τb, the fluctuating torque τf, the spin pumping torque τsp, and the\nspin transfer torque τstt.\nVI. CONCLUSION\nIn this paper, we have presented a general quantum\ntheory of spin dynamics in magnet-normal metal systems,\ngeneralizing earlier results to a general antiferromagnetic\nor ferrimagnetic bipartite lattice. Spin and charge ac-\ncumulations in the normal metals influence the magne-\ntization dynamics in the magnet through spin transfer\ntorque, and the damping is enhanced due to spin pump-\ning, including both inter- and intra-lattice contributions.\nWe derived expressions for transverse fluctuating fields\narising due to the electron magnon interactions. These\nfields have contributions from equilibrium terms as well\nas charge and spin accumulation in the normal metals.\nWe found site-dependent shot noise contributions that\nare non-negligible at low temperatures.ACKNOWLEDGMENTS\nThis work was supported by the Research Council\nof Norway through its Centers of Excellence funding\nscheme, Project No. 262633, ”QuSpin”.\nAppendix A: Holstein-Primakoff transformation\nIn this Appendix, we discuss the transformations used\nto diagonalize the magnon Hamiltonian of Eq. (9). To\ngo from the SU(2) spin operators to bosonic annihila-\ntion and creation operators, we employ the Holstein-\nPrimakoff transformation [37, 38] at sublattices AandB\nand expand to the lowest order in the bosonic operators,\nassuming the antiferromagnet is close to the N´ eel state,\ni.e. that all spins on sublattice A(B) is close to being\nparallel (antiparallel) to the z-direction. At sublattice A,14\nwe expand\nˆSa+=ℏ√\n2sa\u0012\n1−ˆa†\naˆaa\n2sa\u00131/2\nˆaa≈ℏ√\n2saˆaa (A1)\nˆSa−=ℏ√\n2saˆa†\na\u0012\n1−ˆa†\naˆaa\n2sa\u00131/2\n≈ℏ√\n2saˆa†(A2)\nˆSaz=ℏ(sa−ˆa†\naˆaa), (A3)\nwhere aaannihilates a localized magnon and sais the\ntotal spin at lattice site a. In the expansion of the square\nroots in Eq. (A1) and Eq. (A2), we assumed sa≫1 and\nexpanded the square root to lowest order in 1 /sA. We\nhave employed the standard raising and lowering spin\noperators, defined as S±=Sx±iSy.\nSimilarly, at sublattice B, we expand\nˆSb+=ℏ√\n2sbˆb†\nb \n1−ˆb†\nbˆbb\n2sb!\n≈√\n2sbˆb†\nb, (A4)\nˆSb−=ℏ√\n2sb \n1−ˆb†\nbˆbb\n2sb!\nˆbb≈√\n2sbˆbb, (A5)\nˆSbz=ℏ\u0010\n−sb+ˆb†\nbˆbb\u0011\n, (A6)\nwhere ˆbannihilates a localized spin-up magnon.\nAppendix B: Relating spin transfer torque and spin\npumping coefficients\nWe relate the spin transfer pumping coefficients de-\nfined in Eq. (82) to the spin transfer coefficients found\nin Eq. (62) in the case of one normal metal reservoir us-\ning the Onsager reciprocal relations [36]. We start by\ndefining the pumped spin current (in units of electrical\ncurrent, i.e. Ampere) into normal metal as the change in\ntotal spin inside the antiferromagnetic due to spin pump-\ning, i.e.\nIS=−e\nℏX\njSjτsp\nj. (B1)\nThe appearance of Sj=ℏp\nsj(sj+ 1) is due to the way\nwe have defined the torques in the main text, causing\nthem to have the dimension of inverse time. The dynam-\nics of the localized magnetic moment µj=−γSjmjand\nthe spin current are driven by the external effective field\nHeffand the spin accumulation µS, which are the ther-\nmodynamic forces in our system. In linear response, we\ncan then write the equations for the spin dynamics and\nthe spin current in matrix form:\n\u0012−γSi∂tmi\nIS\u0013\n=\u0012Lmm\nijLms\ni\nLsm\njLss\u0013\u0012µ0Heff\nj\nµS/e\u0013\n, (B2)\nwhere the matrix elements 3 ×3 tensors that effectively\napply the relevant cross products to make Eq. (B2) con-\nsistent with the Landau-Lifshitz equation, and where\nwe use the Einstein summation convention for repeated\nLatin indices.1. Identifying Lsm\nInserting the spin pumping torque from Eq. (29), the\nspin current becomes\nIS=−Xj∂tmj, (B3)\nwhere we defined the 3 ×3 matrix Xjas\nXj=e\nℏSjX\nih\nαR\nij˜O+αI\nij˜O2i\n, (B4)\nand the 3 ×3 matrix ˜Oimplements the cross product\nz×v=˜Ovand can be defined in terms of the Levi-\nCivita tensor. The LLG equation in the absence of spin\naccumulation (causing the spin transfer torque to vanish)\nreads\n(1−αb˜O)∂tmi=˜O(−γµ0Heff\ni), (B5)\nwhere αbis the (bulk) Gilbert damping constant. Hence,\nwe identify\nLsm\nj=γXj˜O(1−αb˜O)−1. (B6)\n2. Identifying Lms\nInserting the spin transfer torque from Eq. (27) into\nthe LLGS equation in the absence of an effective field,\nwe find\n∂tmi=ℏ−1(1−αb˜O)−1h\nβI\ni˜O−βR\ni˜O2i\nµS,(B7)\nmeaning that we can identify the linear response coeffi-\ncient Lmsas (no Einstein summation)\nLms\ni=−Siγe\nℏ(1−αb˜O)−1h\nβI\ni˜O−βR\ni˜O2i\n. (B8)\n3. Deriving relations from the Onsager reciprocal\nrelations\nWe are now looking to employ Onsager’s reciprocal\nrelation:\n[Lsm\ni({−mj})]T=Lms\ni({mj}), (B9)\nwhere the superscript Tindicates a matrix transpose in\nthe 3 ×3 Cartesian space. Using the matrix identity\n˜O3=−˜O, we find that Eq. (B9) implies that\nβI\nj˜O−βR\nj˜O2=X\nih\nαI\nij˜O−αR\nij˜O2i\n(B10)\nThis equality is satisfied if\nβj=X\niαij, (B11)15\nwhich generalizes the result from Ref. 35. Inserting the\ndefinitions of these coefficients in the low-temperature\nlimit, we find that\nX\nnWnn\nj↑↓= iπX\ninmWnm\ni↓↑Wmn\nj↑↓, (B12)\nwhich we classify as a generalized optical theorem, since\nin the diagonal case i=j, we can rewrite the imaginary\npart of this to\nIm\"X\nnWnn\ni↑↓#\n=πX\ninm|Wnm\ni↓↑|2, (B13)\nwhich is reminiscent of the optical theorem in wave scat-\ntering theory.\nAppendix C: Contour fields and Keldysh rotations\nIn this Appendix, we show how the action can be writ-\nten in the ±basis, and introduce the Keldysh rotated\nfields, which differ in the case of fermionic and bosonic\nfields. In the ±field basis, the action of the scattering\n(electron) states, corresponding to the Hamiltonian in\nEq. (7), reads\nSe+S0=X\nsZ∞\n−∞dt¯c+\ns(iℏ∂t−ϵ)c+\ns\n−X\nsZ∞\n−∞dt¯c−\ns(iℏ∂t−ϵ)cs−\n=X\nsξt¯cξ\ns(iℏ∂t−ϵ)cξ\ns, (C1)\nwhere now csis a vector containing the scattering fields,\n¯csdenotes its complex conjugate, and ϵis a diagonal\nmatrix containing all energy eigenvalues of the scatter-\ning states. In the final line, we have written the time\nintegration as a sum for concise notation. Additionally,\nwe introduced the sum over ” ±” fields as a sum over\nξ={+,−}, with an implicit negative sign before the ”-”\nfield, i.e.P\nξ. . .ξ=. . .+−. . .−. A similar notation will\nalso be used for the magnon fields below. The negative\nsign ( ξ=−) in the integral in Eq. (C1) and in the other\nactions below originates from reversing the integrationlimits on the backward contour. The magnon action is\nSm=X\nξabt[¯aξ\na(iℏ∂t−EA\nab)aξ\na+¯bξ\nb(iℏ∂t−EB\nab)bξ\nb]\n−2X\naa′Jaa′√sasa′¯aξ\naaξ\na′\n−2X\nbb′Jbb′√sbsb′¯bξ\nbbξ\nb′\n−2X\nξabtJab√sasb[aξ\nabξ\nb+ ¯aξ\na¯bξ\nb]\n−γµ0ℏX\nξatrsa\n2[HA\na−aξ\na+HA\na+¯aξ\na]\n−γµ0ℏX\nξbtrsb\n2[HB\nb−¯bξ\nb+HB\nb+bξ\nb]. (C2)\nThe first-order electron-magnon interaction is\nS1=−X\nξat\nαβr\n2\nsah\naξ\na¯cξ\n↓αWαβ\na↓↑cξ\n↑β+ ¯aξ\na¯cξ\n↑αWαβ\na↑↓cξ\n↓βi\n−X\nξbt\nαβr\n2\nsbh\n¯bξ\nb¯cξ\n↓αWαβ\nb↓↑cξ\n↑β+bξ\nb¯cξ\n↑αWαβ\nb↑↓cξ\n↓βi\n,\n(C3)\nand the second-order term is\nS2=X\nξat\nαβ1\nsa¯aξ\naaξ\nah\n¯cξ\n↑αWαβ\na↑↑cξ\n↑β−¯cξ\n↓αWαβ\na↓↓cξ\n↓βi\n−X\nξbt\nαβ1\nsb¯bξ\nbbξ\nbh\n¯cξ\n↑αWαβ\nb↑↑cξ\n↑β−¯cξ\n↓αWαβ\nb↓↓cξ\n↓βi\n.(C4)\nFor a general bosonic field ϕ, the classical ( cl) and\nquantum ( q) fields are defined as [30]:\nϕcl/q=1√\n2(ϕ+±ϕ−)¯ϕcl/q=1√\n2(¯ϕ+±¯ϕ−).(C5)\nIn our case, we have ϕ={a, b}. The upper (lower) sign\nholds for the classical (quantum) fields. For a fermionic\nfield c, the rotated fields are denoted by 1 and 2, and\ndefined as\nc1/2=1√\n2(c+±c−) ¯c1/2=1√\n2(¯c+∓¯c−).(C6)\nFor fermions, ¯ candcare independent variables, not re-\nlated by complex conjugation.\nAppendix D: Fourier transform\nFor a general function of relative time t−t′, we define\nthe Fourier transform between the relative time domain16\nand the energy domain as\nf(ω) =Z∞\n−∞d(t−t′)eiω(t−t′)f(t−t′), (D1)\nf(t−t′) =Z∞\n−∞dω\n2πe−iω(t−t′)f(ω). (D2)\nThe delta function can be represented as\nδ(t−t′) =Z∞\n−∞dω\n2πe−iω(t−t′), (D3)\nδ(ω) =1\n2πZ∞\n−∞d(t−t′)eiω(t−t′). (D4)Finally, we note a frequently employed identity,\n−iZ∞\n−∞d(t−t′)eiω(t−t′)θ(t−t′) = (ω+ iδ)−1,(D5)\niZ∞\n−∞d(t−t′)eiω(t−t′)θ(t′−t) = (ω−iδ)−1,(D6)\nwhere δis an infinitesimal positive quantity.\n[1] A. Hirohata, K. Yamada, Y. Nakatani, I.-L. Prejbeanu,\nB. Di´ eny, P. Pirro, and B. Hillebrands, Journal of Mag-\nnetism and Magnetic Materials 509, 166711 (2020).\n[2] A. Brataas, B. van Wees, O. Klein, G. de Loubens, and\nM. Viret, Phys. Rep. 885, 1 (2020).\n[3] L. Berger, Phys. Rev. B 54, 9353 (1996).\n[4] J. Slonczewski, Journal of Magnetism and Magnetic Ma-\nterials 159, L1 (1996).\n[5] A. Brataas, Yu. V. Nazarov, and G. E. W. Bauer, Phys-\nical Review Letters 84, 2481 (2000).\n[6] Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, Phys.\nRev. Lett. 88, 117601 (2002).\n[7] B. Heinrich, Y. Tserkovnyak, G. Woltersdorf, A. Brataas,\nR. Urban, and G. E. W. Bauer, Phys. Rev. Lett. 90,\n187601 (2003).\n[8] Y. Tserkovnyak, A. Brataas, G. E. W. Bauer, and B. I.\nHalperin, Rev. Mod. Phys. 77, 1375 (2005).\n[9] J. Foros, A. Brataas, Y. Tserkovnyak, and G. E. W.\nBauer, Phys. Rev. Lett. 95, 016601 (2005).\n[10] A. L. Chudnovskiy, J. Swiebodzinski, and A. Kamenev,\nPhys. Rev. Lett. 101, 066601 (2008).\n[11] J. Swiebodzinski, A. Chudnovskiy, T. Dunn, and\nA. Kamenev, Phys. Rev. B 82, 144404 (2010).\n[12] S. A. Bender, R. A. Duine, and Y. Tserkovnyak, Phys.\nRev. Lett. 108, 246601 (2012).\n[13] S. A. Bender, R. A. Duine, A. Brataas, and\nY. Tserkovnyak, Phys. Rev. B 90, 094409 (2014).\n[14] B. Divinskiy, H. Merbouche, V. E. Demidov, K. O. Niko-\nlaev, L. Soumah, D. Gou´ er´ e, R. Lebrun, V. Cros, J. B.\nYoussef, P. Bortolotti, A. Anane, and S. O. Demokritov,\nNature Communications 12, 6541 (2021).\n[15] L. N´ eel, Magnetism and the local molecular field, Nobel\nLecture (1970).\n[16] T. Jungwirth, X. Marti, P. Wadley, and J. Wunderlich,\nNat. Nanotechnol. 11, 231 (2016).\n[17] S. M. Rezende, A. Azevedo, and R. L. Rodr´ ıguez-Su´ arez,\nJ. Appl. Phys. 126, 151101 (2019).\n[18] A. S. N´ u˜ nez, R. A. Duine, P. Haney, and A. H. MacDon-\nald, Phys. Rev. B 73, 214426 (2006).[19] Z. Wei, A. Sharma, A. S. Nunez, P. M. Haney, R. A.\nDuine, J. Bass, A. H. MacDonald, and M. Tsoi, Phys.\nRev. Lett. 98, 116603 (2007).\n[20] V. Baltz, A. Manchon, M. Tsoi, T. Moriyama, T. Ono,\nand Y. Tserkovnyak, Reviews of Modern Physics 90,\n015005 (2018).\n[21] R. Cheng, J. Xiao, Q. Niu, and A. Brataas, Phys. Rev.\nLett. 113, 057601 (2014).\n[22] P. Vaidya, S. A. Morley, J. van Tol, Y. Liu, R. Cheng,\nA. Brataas, D. Lederman, and E. del Barco, Science 368,\n160 (2020).\n[23] F. Keffer and C. Kittel, Phys. Rev. 85, 329 (1952).\n[24] S. A. Gulbrandsen and A. Brataas, Phys. Rev. B 97,\n054409 (2018).\n[25] Ø. Johansen and A. Brataas, Phys. Rev. B 95, 220408\n(2017).\n[26] J. Barker and O. A. Tretiakov, Physical Review Letters\n116, 147203 (2016).\n[27] A. Kamra and W. Belzig, Phys. Rev. Lett. 119, 197201\n(2017).\n[28] A. Kamra, R. E. Troncoso, W. Belzig, and A. Brataas,\nPhys. Rev. B 98, 184402 (2018).\n[29] J. Tang and R. Cheng, APL Materials 11, 111117 (2023).\n[30] A. Kamenev, Field Theory of Non-Equilibrium Systems\n(Cambridge University Press, Cambridge ; New York,\n2011).\n[31] L. V. Keldysh, Zh. Eksp. Teor. Fiz. (1964).\n[32] S. Takei, Phys. Rev. B 100, 134440 (2019).\n[33] R. A. Duine, A. S. N´ u˜ nez, J. Sinova, and A. H. MacDon-\nald, Physical Review B 75, 214420 (2007).\n[34] C. Tassi, M. Barbieri, and R. Raimondi, Physical Review\nB100, 184418 (2019).\n[35] A. Brataas, Phys. Rev. B 106, 064402 (2022).\n[36] A. Brataas, Y. Tserkovnyak, G. E. W. Bauer, and\nP. J. Kelly, Spin Pumping and Spin Transfer (2012),\narxiv:1108.0385.\n[37] T. Holstein and H. Primakoff, Phys. Rev. 58, 1098\n(1940).\n[38] S. M. Rezende and R. M. White, Physical Review B 14,\n2939 (1976)." }, { "title": "2110.11152v1.The_spin_spin_problem_in_Celestial_Mechanics.pdf", "content": "THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS\nALESSANDRA CELLETTI, JOAN GIMENO, AND MAURICIO MISQUERO\nAbstract. We study the dynamics of two homogeneous rigid ellipsoids subject to\ntheir mutual gravitational in\ruence. We assume that the spin axis of each ellipsoid\ncoincides with its shortest physical axis and is perpendicular to the orbital plane. Due\nto such assumptions, the problem is planar and depends on particular parameters of\nthe ellipsoids, most notably, the equatorial oblateness and the \rattening with respect\nto the shortest physical axes. We consider two models for such con\fguration: while\nin the full model, there is a coupling between the orbital and rotational motions, in\nthe Keplerian model, the centers of mass of the bodies are constrained to move on\ncoplanar Keplerian ellipses. The Keplerian case, in the approximation that includes\nthe coupling between the spins of the two ellipsoids, is what we call spin-spin problem,\nthat is a generalization of the classical spin-orbit problem. In this paper we continue the\ninvestigations of [Mis21] on the spin-spin problem by comparing it with the spin-orbit\nproblem and also with the full model.\nBeside detailing the models associated to the spin-orbit and spin-spin problems, we\nintroduce the notions of standard and balanced resonances, which lead us to investi-\ngate the existence of periodic and quasi-periodic solutions. We also give a qualitative\ndescription of the phase space and provide results on the linear stability of solutions for\nthe spin-orbit and spin-spin problems. We conclude by providing a comparison between\nthe full and the Keplerian models with particular reference to the interaction between\nthe rotational and orbital motions.\n1.Introduction\nThe dynamics of two rigid bodies orbiting under their mutual gravitational attraction is\na classical problem of Celestial Mechanics known as the Full Two-Body problem . In this\ncontext, Kinoshita investigated the problem by using Hori-Deprit perturbation theory\n[Kin72], assuming that one of the bodies is spherical and the other body is triaxial. Later,\nthe problem of two extended rigid bodies was studied in [Mac95] as a Hamiltonian system\nwith respect to a non-canonical structure, which is used to characterize the relative\nequilibria. A seminal work was performed in [Bou17] to which we refer for an alternative\ndescription of the model of the full two rigid body problem using spherical harmonics\nand Wigner D-matrices. In [Sch09], the problem is restricted to a planar con\fguration\nwith the potential expanded to order 1 =r3, whereris the relative distance between the\ntwo rigid bodies; under this condition, [Sch09] describes the relative equilibria and their\nstability properties.\nKey words and phrases. Spin-spin model jSpin-orbit model jTwo-body problem jResonancesj\nPeriodic orbitsjQuasi-periodic solutions.\nA.C. and M.M. thank MSCA-ITN-ETN Stardust-R, Grant Agreement 813644 and J.G. thanks MIUR-\nPRIN 20178CJA2B \\New Frontiers of Celestial Mechanics: theory and Applications\". J.G. has also been\nsupported by the Spanish grants PGC2018-100699-B-I00 (MCIU/AEI/FEDER, UE) and the Catalan\ngrant 2017 SGR 1374.\n1arXiv:2110.11152v1 [math.DS] 21 Oct 20212 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nIn this paper, we investigate di\u000berent simpli\fed models of rotational dynamics of\ncelestial bodies, subject to the mutual gravitational attraction. The spin-spin problem\nwas introduced in [Mis21] as a planar version of the Full Two-Body problem for ellipsoids\n(compare with [BM15, BL09]), by using the expansion of the potential up to order 1 =r5,\nwhich results in the coupling of the spins of both bodies. An equivalent model was\nstudied in [JA16] (see also [HX17]).\nIndeed, we consider a hierarchy of models with di\u000berent complexity. In particular, we\nstart by considering two homogeneous rigid ellipsoidal bodies subject to the following\nassumptions:\nAssumption 1. The spin axis of each ellipsoid is perpendicular to the orbital plane.\nAssumption 2. The spin axis of each ellipsoid is aligned with the shortest physical axis\nof the satellite.\nAssumptions 1 and 2 imply that the motion takes place on a plane. Following [Mis21],\nwe introduce a Hamiltonian function that includes both the orbital and rotational mo-\ntions. Using the conservation of the angular momentum, the system is described by a\nHamiltonian with 3 degrees of freedom that depends on several parameters of each ellip-\nsoid, among which there are the equatorial oblateness and the \rattening with respect to\nthe shortest physical axis. Such parameters are typically small for natural bodies of the\nsolar system. We refer to this model as the fullproblem, since it includes the coupling\nbetween the orbital and rotational motions.\nThe potential of the problem can be written as V=V0+P1\nl=1V2l, whereV0denotes\nthe Keplerian potential and the terms V2lare proportional to 1 =r2l+1, whereris the\ninstantaneous distance of the two centers of mass. If we consider the expansion up to\norderl= 2, sayV=V0+V2+V4, we obtain that the model includes the coupling of\nthe spins of the two ellipsoids through the term V4, which contains combinations of the\nrotation angles of the two satellites. When the two spins interact, we refer to the problem\nas the full spin-spin model. If, instead, we limit the potential to V=V0+V2, then we\nobtain two decoupled systems, the full spin-orbit models (see, e.g., [Bel66, Cel90, GP66]).\nWhen one of the bodies is spherical, its spin is uniform and the dynamics of the spin-spin\nmodel becomes very similar to that of the spin-orbit model, but including the terms in\nV4.\nNext, we introduce another assumption, namely:\nAssumption 3. The orbital motion of the ellipsoids coincides with that of two point\nmasses, so that both centers of mass move on coplanar Keplerian orbits with eccentricity\ne2[0;1) and with a common focus at the barycenter of the system.\nAssumption 3 implies that the orbit is not a\u000bected by the rotational motion. To\nthefullspin-spin and spin-orbit problems, it corresponds the Keplerian spin-spin and\nspin-orbit models, described by a Hamiltonian function with an explicit periodic time\ndependence.\nNote that we are considering rigid bodies only, which means that dissipative e\u000bects\ndue to tidal torques are not considered. We refer to [CC08, CC09, MO20, Mis21, GP66]\nfor a description of the dissipative spin-orbit and spin-spin problems.THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 3\nThe previously described models have some symmetries that are a direct consequence\nof the mirror symmetries of the ellipsoids. This fact leads us to introduce the following\ntwo types of resonances within the spin-orbit problem of a single ellipsoid:\n(R1) We call standardm:nspin-orbit resonance, for some integers m,n, when the\nspinning body makes mrotations during norbital revolutions;\n(R2) We call balancedm: 2 spin-orbit resonance, for some integers m, when the\nspinning body makes m=2 of a rotation during one orbital revolution.\nWe remark that a balanced spin-orbit resonance is also a standard resonance, but the\nconverse is not always true. For example, a balanced 2 k: 2 resonance for k2Zis\nequivalent to a k: 1 spin-orbit resonance, but there is not such an equivalence for order\n(2k+ 1) : 2. Both de\fnitions ( R1) and ( R2) extend to the spin-spin problem of type\n(m1:n1;m2:n2) for integers m1,m2,n1,n2, when the \frst ellipsoid is in a m1:n1\nspin-orbit resonance and the second ellipsoid in a m2:n2spin-orbit resonance.\nWe also stress that spin-orbit resonances \fnd many applications in the solar system;\nin fact, the Moon is an example of a 1 : 1 spin-orbit resonance1, since it makes a rotation\nin the same period it takes to make an orbit around the Earth. This is also called a\nsynchronous spin-orbit resonance, which is common to many satellites of other planets,\nincluding Mars, Jupiter, Saturn, Uranus, Neptune. Among the planets, Mercury is locked\nin a 3 : 2 spin-orbit resonance around the Sun. On the other hand, the Pluto-Charon\nsystem is locked in the double synchronous spin-spin resonance (1 : 1 ;1 : 1).\nIn this work, we study the behavior of the solutions of the spin-orbit and spin-spin\nproblems as the parameters and the initial conditions are varied. In particular, we\ninvestigate the boundary conditions that lead to the existence of symmetric periodic\norbits. Such results (see Propositions 5 and 10) use some symmetry properties of the\nequations of motion. We remark that these symmetries are lost if we include dissipation;\nhowever, such periodic solutions might be continued to the dissipative setting as shown\nin [MO20, Mis21]. Beside the study of the periodic orbits, we provide the conditions for\nthe existence of quasi-periodic solutions of the Keplerian version of the spin-spin model.\nWe also give a qualitative study of the spin-orbit problem as well as the spin-spin\nproblem with spherical and non-spherical companion. Within such investigation, we\ndiscover some new features, like the measure synchronization (see, e.g., [HZ99]) for the\nspin-spin problem with identical bodies. Our study leads to analyze the multiplicity\nof solutions and the linear stability of the periodic orbits (compare with [CC00]). In\ngeneral, we \fnd that there is not a unique solution associated to a particular resonance;\nhowever, for some values of the parameters such a uniqueness exists. Finally, we provide\nsome results on the comparison between the full and Keplerian models, as well as on\nthe interaction between the spin and the orbital motion, motivated by the fact that the\ncoupling between the rotational and orbital motions has not been much explored in the\nliterature.\nThis work is organized as follows. In Section 2 we present the spin-orbit and spin-spin\nmodels. The de\fnition of resonances and the existence of periodic and quasi-periodic\norbits are given in Section 3. A qualitative description of the phase space is given in\n1A 2 : 2 balanced spin-orbit resonance is equivalent to a 1 : 1 standard spin-orbit resonance.4 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nSection 4. The linear stability of symmetric periodic orbits is investigated in Section 5.\nFinally, a comparison between the full and Keplerian models is presented in Section 6.\n2.The models for the spin-orbit and spin-spin coupling\nThe aim of this section is to present the so-called spin-orbit andspin-spin models, that\nwe are going to introduce as follows. The assumptions and the notations are given in\nSection 2.1; di\u000berent models, subject to some or all the assumptions listed in Section 2.1,\nare presented in Sections 2.2 and 2.3.\n2.1.General assumptions. Consider two homogeneous rigid ellipsoids, say E1andE2,\nwith masses M1andM2, respectively. Let Ajbj>cj. We refer to the\nFull Two-Body Problem (hereafter F2BP) as the problem of two rigid bodies interacting\ngravitationally (see, e.g., [Sch02, Sch09]). When the bodies have ellipsoidal shape, we\nspeak of the ellipsoidal F2BP , where we make the assumptions 1, 2, 3 of Section 1:\norbit of\norbit of\nFigure 1. The planar spin-spin problem considering Assumptions 1 to 3.\nAssumptions 1 and 2 guarantee that the problem we deal with is a planar problem.\nAdditionally, Assumption 3 restricts the problem so that we obtain a model with two\ndegrees of freedom and a periodic time dependence. This assumption is equivalent to\nsay that the spin motion will not in\ruence the orbital motion. Besides, note that we\nare neglecting the gravitational contribution of other bodies, we are not considering any\ndissipative e\u000bect that might arise, for example, from the non-rigidity of the ellipsoidal\nbodies and we do not take into account the obliquity, namely the inclination of the\nspin-axes with respect to the orbital plane.\nWe are going to work with units adapted to the system. If we call \u001cthe orbital period\nof the Keplerian orbit, then we will use units such that\nM1+M2= 1;C1+C2= 1; \u001c = 2\u0019 :\nRecall Kepler's third law for the Two-Body Problem\nG(M1+M2)\u0010\u001c\n2\u0019\u00112\n=a3;THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 5\nwhereGis the gravitational constant, ais the semi-major axis associated to the motion\nof the reduced mass of the system, say \u0016=M1M2in our units. In consequence, G=a3\nin our units.\nLet us now de\fne the parameters for each ellipsoid\ndj=Bj\u0000Aj; qj= 2Cj\u0000Bj\u0000Aj;\nthe quantity dj=Cjmeasures the equatorial oblateness of each ellipsoid with respect to\nthe plane formed by the directions of ajand bj, whereasqj=Cjmeasures the \rattening\nwith respect to the direction corresponding to the cj-axis.\nNote that ifAj\u0014Bj\u0014Cj, then, in our units there are some bounds for the parameters\nof the system given by\n0\u0014dj\u0014Cj\u00141; dj\u0014qj\u00142Cj\u00142; Mja2\nj=5\n2(Cj+dj)\u00145Cj\u00145: (1)\nThe last relation in (1) comes from the fact that the moments of inertia of an ellipsoid\nhold the identities\nAj=1\n5Mj(b2\nj+c2\nj);Bj=1\n5Mj(a2\nj+c2\nj);Cj=1\n5Mj(a2\nj+b2\nj):\n2.2.The full models. First, let us derive the equations of the full models of spin-orbit\nand spin-spin coupling, for which only Assumptions 1 and 2 hold, namely we do not\nconstrain the centers of mass of E1andE2to move on Keplerian ellipses.\nThe equations of motion are obtained by computing the Hamiltonian function through\na Legendre transformation of the Lagrangian, say L=T\u0000V, whereTis the kinetic\nenergy and Vthe potential energy of the system. We split Tin two parts, associated\nrespectively to the orbital and rotational motions, say T=Torb+Trot.\nLet us identify the orbital plane with the complex plane C, consider the center of mass\nof the system \fxed in the origin and let the position of each ellipsoid be rj2C. Then,\nby de\fnition of the barycenter we have\nM1r1+M2r2=0:\nIf we de\fne the relative position vector r=r2\u0000r1, since in our units M1+M2= 1, then\nwe have\nr1=\u0000M2r;r2=M1r: (2)\nFigure 2. Generalized coordinates of the full planar model.6 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nWe introduce the Lagrangian generalized coordinates ( r;f;\u0012 1;\u00122), illustrated in Fig-\nure 2, in the following way. The distance r > 0 and the angle fde\fne the relative\nposition vector r=rexp(if)2Cwith respect to an inertial reference frame with a \fxed\nx-axis. The angles \u00121,\u00122provide the orientation of the axes a1,a2with respect to the\nx-axis. The variables randfde\fne the orbital motion of the system and \u0012jthe spin\nmotion of the ellipsoid Ej.\nTo write the kinetic energy, we notice that we can express rin components as r=\n(rcosf;rsinf), which gives _r= ( _rcosf\u0000r_fsinf;_rsinf+r_fcosf). Then, using (2),\nthe total orbital and the rotational kinetic energies are given by\nTorb=1\n2(M1_r2\n1+M2_r2\n2) =\u0016\n2_r2=\u0016\n2( _r2+r2_f2); Trot=1\n2C1_\u00122\n1+1\n2C2_\u00122\n2;\nwhere\u0016=M1M2in our units. The Lagrangian is given by\nL(r;f;\u0012 1;\u00122;_r;_f;_\u00121;_\u00122) =Torb(r;_r;_f) +Trot(_\u00121;_\u00122)\u0000V(r;f;\u0012 1;\u00122);\nwhere, according to [Mis21, Bou17], the full expansion of the potential energy of the\nsystemV=V(r;f;\u0012 1;\u00122) takes the form\nV(r;f;\u0012 1;\u00122) =\u0000GM 1M2\nrX\n(l1;m1)2\u0007\n(l2;m2)2\u0007\u0003l1;m1\nl2;m2\nr2(l1+l2)cos(2m1(\u00121\u0000f) + 2m2(\u00122\u0000f)); (3)\nwhere\n\u0007 =f(l;m)2Z2: 0\u0014jmj\u0014lg;\nand the constants \u0003l1;m1\nl2;m2are de\fned in Appendix A; we refer to [Mis21] for full details.\nLet us de\fne the momenta as\npr=@_rL=\u0016_r; pf=@_fL=\u0016r2_f; pj=@_\u0012jL=Cj_\u0012j;\nthen, the Hamiltonian of the system becomes\nH(r;f;\u0012 1;\u00122;pr;pf;p1;p2) =p2\nr\n2\u0016+p2\nf\n2\u0016r2+p2\n1\n2C1+p2\n2\n2C2+V(r;f;\u0012 1;\u00122): (4)\nConsequently, the equations of motion are\n_r=pr\n\u0016;_f=pf\n\u0016r2;_pr=p2\nf\n\u0016r3\u0000@rV; _pf=\u0000@fV (5)\nand\n_\u0012j=pj\nCj;_pj=\u0000@\u0012jV : (6)\nWe remark that with these variables, the problem splits in two parts: equations (5)\ndescribe the orbital motion, while equations (6) describe the rotational motion. The\nevolution of both parts is coupled through the potential V.\nThe potential V, expanded in (3), can be written in a perturbative way as\nV(r;f;\u0012 1;\u00122) =V0(r) +Vper(r;f;\u0012 1;\u00122); V 0(r) =\u0000GM 1M2\nr: (7)\nWe remark that the term V0corresponds to the classical Keplerian form for the poten-\ntial and the term Vperprovides the coupling between the spin and the orbital motions.THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 7\nMoreover, we can expand VperasVper=P1\nl=1V2l;whereV2lare suitable terms propor-\ntional to 1=r2l+1. A truncation of the expansion of Vperwill result in an approximated\ndynamics of our system. The explicit expressions of the \frst two terms of such a expan-\nsion are given in [Mis21] and we report them here:\nV2=\u0000GM 2\n4r3(q1+ 3d1cos(2\u00121\u00002f))\u0000GM 1\n4r3(q2+ 3d2cos(2\u00122\u00002f));\nV4=\u00003G\n43r5(\n12q1q2+15\n7[M2\nM1d2\n1+ 2M2\nM1q2\n1+M1\nM2d2\n2+ 2M1\nM2q2\n2]\n+d1M2\u001a\n[20q2\nM2+100\n7q1\nM1] cos(2\u00121\u00002f) + 25d1\nM1cos(4\u00121\u00004f)\u001b\n+d2M1\u001a\n[20q1\nM1+100\n7q2\nM2] cos(2\u00122\u00002f) + 25d2\nM2cos(4\u00122\u00004f)\u001b\n+6d1d2cos(2\u00121\u00002\u00122) + 70d1d2cos(2\u00121+ 2\u00122\u00004f))\n: (8)\nFrom now on, we will refer to (5) and (6) as the full models: full spin-orbit model if we\ntakeVper=V2, in which the angles \u00121,\u00122appear in di\u000berent trigonometric terms, and full\nspin-spin model ifVper=V2+V4, which contains trigonometric terms with combinations\nof the rotation angles \u00121,\u00122. These names are motivated by the well-known spin-orbit\nmodel, [Cel10], and the spin-spin model from [Mis21]. If we consider the models under\nAssumption 3 which gives a constraint on the orbit, we speak of Keplerian spin-orbit\nmodel andKeplerian spin-spin model (compare with Section 2.3).\n2.2.1. Conservation of the angular momentum. Note that the Hamiltonian Hin (4) is\ninvariant under the transformation ( r;f;\u0012 1;\u00122)7!(r;f+\u000ef;\u0012 1+\u000ef;\u0012 2+\u000ef), where\u000ef\nis an in\fnitesimal angular increase, because the angular arguments of V(r;f;\u0012 1;\u00122) only\ndepend on the di\u000berences \u00121\u0000fand\u00122\u0000f. This symmetry is related, by Noether's\ntheorem [Gol80], with a conserved quantity, say, the total angular momentum pf+p1+p2.\nThis can be proved through the following change of variables\n(r;f;\u0012 1;\u00122;pr;pf;p1;p2)7!(r;f;\u001e 1;\u001e2;pr;Pf;p1;p2);\nwhere\n\u001ej=\u001ej(f;\u0012j) =\u0012j\u0000f; Pf=Pf(pf;p1;p2) =pf+p1+p2: (9)\nThe transformation of coordinates (9) is canonical, since\ndr^dpr+ df^dpf+2X\nj=1d\u0012j^dpj= dr^dpr+ df^dPf+2X\nj=1d\u001ej^dpj:\nThen, the Hamiltonian (4) in this new set of variables is given by\nH(r;f;\u001e 1;\u001e2;pr;Pf;p1;p2) =p2\nr\n2\u0016+(Pf\u0000p1\u0000p2)2\n2\u0016r2+p2\n1\n2C1+p2\n2\n2C2+V(r;\u001e1;\u001e2);(10)8 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nwhereV(r;\u001e1;\u001e2) =V(r;f;\u001e 1+f;\u001e 2+f). Now it is clear that fis an ignorable variable\nin (10) and that Pfis a constant of motion, corresponding to the total angular momentum\nof the system.\nIn summary, in the evolution of the system, there is a transfer of angular momentum\nbetween the spin part, given for each body by pj=Cj_\u0012j, and the orbital part, given by\npf=\u0016r2_f.\n2.3.The Keplerian models. In this section, we introduce the Assumption 3 to the\nmodel of Section 2.2. From (5) and (6), it is straightforward to constrain the orbit\nto be Keplerian; only in the orbital part we retain just the term V0of the potential\nV=V0+Vperin (7), thus obtaining the following equations:\n_r=pr\n\u0016;_f=pf\n\u0016r2;_pr=p2\nf\n\u0016r3\u0000@rV0;_pf=\u0000@fV0= 0; (11)\nand\n_\u0012j=pj\nCj;_pj=\u0000@\u0012jV=\u0000@\u0012jVper: (12)\nA convenient procedure to numerically integrate the equations of motion (12) is pre-\nsented in Appendix B.\n2.3.1. Orbital motion. Note that since @\u0012jV0= 0, the system (11) is now decoupled from\n(12). Moreover, (11) is the Kepler problem, whose solutions depend on the eccentricity\neand the semi-major axis aof the orbit. Here we assume for simplicity that the orbit\nis a 2\u0019-periodic Keplerian ellipse of eccentricity e2[0;1) with focus at the origin and\nwith the periapsis on the positive x-axis.\nSince the orbital period is 2 \u0019, then, we can take the time tto coincide with the mean\nanomaly . We denote by utheeccentric anomaly , which, in our units, is related to the\nmean anomaly by Kepler's equation\nt=u\u0000esinu : (13)\nThe orbital radius is related to uby\nr=a(1\u0000ecosu): (14)\nWe can write the vector r2Cin terms of the eccentric anomaly also as\nrexp(if) =a(cosu\u0000e+ip\n1\u0000e2sinu): (15)\nNote that for t= 0 we assumed, without loss of generality, that f=u= 0, and\nconsequently, f=u=\u0019whent=\u0019. From (15) we obtain the following useful relations\nbetweenfandu\ncosf=cosu\u0000e\n1\u0000ecosu; sinf=p\n1\u0000e2sinu\n1\u0000ecosu: (16)\nWith the previous de\fnitions, the Keplerian orbit of eccentricity eand semi-major\naxisais given by the functions\nr=r(t;a;e); f =f(t;e); pr=pr(t;a;e); pf=pf(a;e) =\u0016a2p\n1\u0000e2; (17)\nthat correspond to the solution of equations (11) generated by the initial conditions\nr(0) =a(1\u0000e); pr(0) = 0; f(0) = 0; pf(0) =\u0016a2p\n1\u0000e2: (18)THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 9\n2.3.2. Spin motion. The spin motion is described by equations (12) with the Keplerian\nperiodic input (17) given implicitly by eqs. (13) to (15). This motion can be described\nby the non-autonomous Hamiltonian\nHK(t;\u00121;\u00122;p1;p2) =H(r(t;a;e);f(t;e);\u00121;\u00122;pr(t;a;e);pf(a;e);p1;p2); (19)\nwhereH(r;f;\u0012 1;\u00122;pr;pf;p1;p2) is the Hamiltonian of the full model de\fned in (4). The\nHamiltonian (19) is hence of the form\nHK(t;\u00121;\u00122;p1;p2) =p2\n1\n2C1+p2\n2\n2C2+W(t;\u00121;\u00122); (20)\nwhere the potential Wis 2\u0019-periodic in tand\u0019-periodic in \u00121and\u00122. The equations of\nmotion (12) take the form\n_\u0012j=pj\nCj;_pj=\u0000@\u0012jW(t;\u00121;\u00122): (21)\nLet us de\fne the non-dimensional parameters of the model:\n\u0015j= 3\u0016\nMjdj\nCj; \u001bj=1\n3Cj\n\u0016a2; ^qj=qj\nMja2; (22)\nwhere\u0015jrepresents the equatorial oblateness of Ej;\u001bjis the ratio between the moment\nof inertia ofEjand the orbital one; and ^ qjmeasures the \rattening of Ejwith respect\nto the size of the orbit. Note that the parameters in (22) are small for bodies that are\nclose to spherical. Besides, not all the parameters de\fned previously are free, because\nwe have the constraint C1\u001b2=C2\u001b1.\nIf we takeVper=V2, then the system (12) becomes\n\u0012j+\u0015j\n2\u0012a\nr(t;e)\u00133\nsin(2\u0012j\u00002f(t;e)) = 0; j = 1;2; (23)\nthat is a system of two uncoupled spin-orbit problems. Each of these problems depends\njust on two parameters: ( e;\u0015j). On the other hand, if Vper=V2+V4, from (12) and (8)\nwe obtain the following system for j= 1;2,\n0 =\u0012j+\u0015j\n2(\u0012a\nr(t;e)\u00133\nsin(2\u0012j\u00002f(t;e))+\n+\u0012a\nr(t;e)\u00135\"\n5\n4\u0012\n^q3\u0000j+5\n7^qj\u0013\nsin(2\u0012j\u00002f(t;e)) +25\n8\u0015j\u001bjsin(4\u0012j\u00004f(t;e))\n+\u00153\u0000j\u001b3\u0000j\u00123\n8sin(2\u0012j\u00002\u00123\u0000j) +35\n8sin(2\u00123\u0000j+ 2\u0012j\u00004f(t;e))\u0013#)\n;(24)\nthat we call spin-spin problem. From the previous discussion, this model depends on\nseven independent parameters2(e;C1;\u00151;\u00152;\u001b1;^q1;^q2). Note that in (24), the coupling\nbetween the dynamics of \u00121and\u00122is given by \u001b1and\u001b2. Moreover, if ^ qj=\u001bj= 0, the\nspin-spin problem (24) is reduced to a pair of spin-orbit problems (23).\n2In [Mis21] there was an error because ^ q1and ^q2are actually independent, it is not always true that\nC1\u00151^q2=C2\u00152^q1, but it can be regarded as an additional constraint.10 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nLet us now consider (24) in the case that E2is a sphere, that is, d2=q2= 0. Then,\n\u00152=\u001b2= ^q2= 0, which implies that E2is in uniform rotation \u00122(t) = _\u00122(0)t+\u00122(0).\nThe dynamics of \u00121is uncoupled from \u00122and is given by\n0 =\u00121+\u00151\n2n\u0012a\nr(t;e)\u00133\nsin(2\u00121\u00002f(t;e))+\n+\u0012a\nr(t;e)\u00135h25^q1\n28sin(2\u00121\u00002f(t;e)) +25\u00151\u001b1\n8sin(4\u00121\u00004f(t;e))io\n;(25)\nthat is a spin-orbit problem up to order 1 =r5. An equivalent system was studied previ-\nously in [JNA16]. Here the parameters \u001b1and ^q1perturb the framework of the spin-orbit\nproblem (23).\n2.3.3. Reversing symmetries. The equations for the spin motion in the Keplerian mod-\nels have some reversing symmetries, i.e., transformations in the phase space that keep\ninvariant the equations of motion with a time reversal.\nDe\fnition 1. Consider the di\u000berential equation\ndx\ndt=F(x);x2Rn; (26)\nand let x(t;x0) be the solution of (26) with initial condition x(0;x0) =x0.\n(1) A transformation R:Rn!Rnis called a reversing symmetry of (26) if\ndR(x)\ndt=\u0000F(R(x)):\n(2) The \fxed point set of Ris given by Fix( R) =fx2Rn:R(x) =xg.\n(3) An orbit o(x0) =fx(t;x0):t2RgisR-symmetric if R(o(x0)) =o(x0).\nThe system (23) with j= 1 can be written in the autonomous form (26) with\nx= (t;\u00121;_\u00121);F(x) = \n1;_\u00121;\u0000\u00151\n2\u0012a\nr(t;e)\u00133\nsin(2\u00121\u00002f(t;e))!\n: (27)\nOne can easily check that each transformation de\fned by\nR\u000b;\f(x) = (2\u000b\u0000t;2\f\u0000\u00121;_\u00121);with (\u000b;\f)2\u0019Z\u0002\u0019\n2Z; (28)\nis a reversing symmetry of equation (26) with Fas in (27) because\nf(2\u000b\u0000t;e) = 2\u000b\u0000f(t;e); r(2\u000b\u0000t;e) =r(t;e):\nThe same is true replacing in (28) the quantities \u00121,_\u00121by\u00122,_\u00122.\nOn the other hand, the spin-spin problem in its Hamiltonian formulation is given by\n(21), that can be written as (26) with\nx= (t;\u00121;\u00122;p1;p2);F(x) =\u0012\n1;p1\nC1;p2\nC2;\u0000@\u00121W(t;\u00121;\u00122);\u0000@\u00122W(t;\u00121;\u00122)\u0013\n:(29)\nEach transformation de\fned by\nR\u000b;\f1;\f2(x) = (2\u000b\u0000t;2\f1\u0000\u00121;2\f2\u0000\u00122;p1;p2);with (\u000b;\f 1;\f2)2\u0019Z\u0002\u0019\n2Z\u0002\u0019\n2Z;(30)\nis a reversing symmetry of equation (26).THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 11\nIn the following sections we are going to emphasize the study of orbits that are sym-\nmetric with respect to the mentioned reversing symmetries using the following lemma.\nLemma 2 (from Theorem 4.1 in [LR98]) .An orbit of (26) is R-symmetric if, and only\nif, it intersects the \fxed point set Fix(R).\n3.Periodic and quasi-periodic solutions of the Keplerian models\nIn this section we provide the de\fnitions of spin-orbit and spin-spin resonances, giving\nsome results on the existence of periodic orbits (Section 3.1) and KAM tori (Section 3.2).\n3.1.The spin-orbit and spin-spin resonances. The periodic solutions of the Kep-\nlerian models presented in the Section 2.3 correspond to resonances between the orbital\nmotion and the spin motion. The expansion of the potential in (8) contains much inter-\nesting information concerning such resonances. First, let us introduce the de\fnition of\nspin-orbit resonances.\nDe\fnition 3. We say that the ellipsoid E1is in a standard spin-orbit resonance of order\nm:nwithm2Z;n2Znf0g, if\n\u00121(t+ 2\u0019n) =\u00121(t) + 2\u0019m : (31)\nThe associated resonant angle m:n\n1(t) =mt\u0000n\u00121(t) is a periodic function of period 2 \u0019n.\nThe same de\fnition holds for E2.\nRecalling that we have normalized the mean motion to unity, we remark that De\fni-\ntion 3 states that the ratio of the orbital period of Ejover its period of rotation is m=n,\nn6= 0. Additionally, according to De\fnition 3, the resonance m:n,n6= 0, is also of\norderkm:kn,k2Znf0g, but the converse is not true in general. For example, with\n(31), the resonance 1 : 1 is also of order 2 : 2, but a resonance 2 : 2 may not be of order\n1 : 1. We can say that the resonance m:nis of higher order than the resonance m0:n0\nifm=n>m0=n0. This will be denoted using the notation m:n>m0:n0.\nNext, we introduce a di\u000berent de\fnition of spin-orbit resonance.\nDe\fnition 4. We say that the ellipsoid E1is in a balanced spin-orbit resonance of order\nm: 2,m2Z, if\n\u00121(t+ 2\u0019) =\u00121(t) +m\u0019 : (32)\nIn this case, the resonant angle m:2\n1(t) is 2\u0019-periodic. The same de\fnition holds for E2.\nNotice that the two notions (31) and (32) are not equivalent: (32) implies (31) for\nm: 2, but the converse is not true. Actually, note that a balanced 2 k: 2 resonance, with\nk2Z, is a spin-orbit resonance of order k: 1. This new de\fnition was motivated by\n[BL75], where the solutions associated to the resonance 3 : 2 of the spin-orbit problem\n(say, (23) with j= 1) were studied numerically. Basically, they found out that the\nsolutions satisfying (31), but not (32), appear only for large values of \u00151(&1), also, for\na given point ( e;\u00151) the solutions appear in multiplets, and \fnally, the corresponding\nresonant angles have large amplitudes ( j 3:2\n1(t)j&0:75), see Table 1 in [BL75]. On the\nother hand, solutions that obey (32) exist for any point in the ( e;\u00151)-plane, including\nlarge regions of uniqueness of solution and resonant angle with small amplitude. Note\nthat such amplitude is a measure of the deviation of the solution with respect to the\nuniform rotation of angular velocity3\n2t. In De\fnition 4 we generalize these two types of12 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nresonances for any order m: 2, since this let us determine the main resonances in the\n\frst orbital revolution.\nFrom [DeV58], for instance, we know that a good tool to study a di\u000berential equation\nthat has some reversing symmetries is by studying the periodic orbits that are invariant\nunder such transformations.\nIn the next proposition we provide some boundary conditions that characterize the\nsymmetric orbits in spin-orbit resonances.\nProposition 5. The following statements hold for the spin-orbit problem (23), with\nj= 1 (ellipsoidE1):\n(1)AnyR\u000b;\f-symmetric orbit, with R\u000b;\fde\fned in (28), associated to a m:nspin-\norbit resonance is equivalent to a solution that satis\fes the following Dirichlet\nconditions:\n\u00121(\u000b) =\f; \u0012 1(\u000b+n\u0019) =\f+m\u0019 ; (33)\nwith\u000b2f0;\u0019gand\f2f0;\u0019\n2g(four combinations). Moreover, such solution\nsatis\fes the following symmetry property \u00121(t) = 2\f\u0000\u00121(2\u000b\u0000t).\n(2)There are two independent types of R\u000b;\f-symmetric orbits representing a balanced\nm: 2spin-orbit resonance and are given by:\nType 0 :\u00121(0) = 0; \u0012 1(\u0019) =m\u0019\n2; (34)\nType\u0019\n2:\u00121(0) =\u0019\n2; \u0012 1(\u0019) =(m+ 1)\u0019\n2: (35)\nMoreover, the corresponding symmetry relations are: \u00121(t) =\u0000\u00121(\u0000t)for type I\nand\u00121(t) =\u0019\u0000\u00121(\u0000t)for type II.\nThe same is true for (23), with j= 2 (ellipsoidE2), and also for (25).\nProof. Let us apply Lemma 2 to (26) with F(x) given by (27). The \fxed point set of each\nreversing symmetry R\u000b;\fis Fix(R\u000b;\f) =f(t;\u00121;_\u00121):t=\u000b;\u0012 1=\fg, then, the symmetric\norbits can be found with initial conditions \u00121(\u000b) =\f. Since F(x) is 2\u0019-periodic in tand\n\u0019-periodic in \u00121, it is enough to consider \u000b2f0;\u0019g(the periapsis and the apoapsis) and\n\f2f0;\u0019\n2g.\nNow, since R\u000b;\fis a reversing symmetry, if \u00121(t) is a solution of (23) with j= 1, so it\nis (t) = 2\f\u0000\u00121(2\u000b\u0000t). Additionally, if \u00121(\u000b) =\f, then, both solutions coincide, so\nthe symmetry relation \u00121(t) = 2\f\u0000\u00121(2\u000b\u0000t) holds for it. Now, if \u00121(t) is in am:n\nspin-orbit resonance, then, replacing t=\u000b\u0000n\u0019in (31), we get\n\u00121(\u000b+n\u0019) =\u00121(\u000b\u0000n\u0019) + 2\u0019m :\nFrom the symmetry relation we get additionally that\n\u00121(\u000b\u0000n\u0019) = 2\f\u0000\u00121(\u000b+n\u0019):\nCombining the two expressions we prove (33).\nNow let us prove the converse, that a solution \u00121(t) satisfying the conditions (33) is in\nm:nspin-orbit resonance. Let the initial conditions of such solution be\n\u00121(\u000b+n\u0019) =\f+m\u0019; _\u00121(\u000b+n\u0019) =~\f ; (36)THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 13\nfor some real constant ~\f. Since\u00121(\u000b) =\f, the solution has the symmetry relations\n\u00121(t) = 2\f\u0000\u00121(2\u000b\u0000t) and _\u00121(t) =_\u00121(2\u000b\u0000t). Using such relations we get that\n\u00121(\u000b\u0000n\u0019) =\f\u0000m\u0019; _\u00121(\u000b\u0000n\u0019) =~\f : (37)\nThe initial conditions (36) and (37) show that, while the time increases 2 \u0019n, the angle\u00121\nincreases in 2 \u0019m, and the angular velocity is the same for both cases, then, the solution\nis periodic. With this, we have proved item 1.\nLet us now consider the balanced m: 2 case in item 2. Note that, using the de\fnition\n(32) instead of (31), we can follow the same procedure as before to prove that a balanced\nm: 2 solution satis\fes the following boundary conditions\n\u00121(\u000b) =\f; \u0012 1(\u000b\u0006\u0019) =\f\u0006m\u0019\n2; (38)\nand the symmetry relation \u00121(t) = 2\f\u0000\u00121(2\u000b\u0000t). Then, we get that the four types are\nin this case (34), with \u000b= 0;\f= 0; (35), with \u000b= 0;\f=\u0019=2;\nType 00:\u00121(0) =\u0000m\u0019\n2; \u0012 1(\u0019) = 0;\nwith\u000b=\u0019;\f= 0, and\nType\u0019\n20\n:\u00121(0) =(1\u0000m)\u0019\n2; \u0012 1(\u0019) =\u0019\n2;\nwith\u000b=\u0019;\f =\u0019=2. Since an ellipsoid has a mirror symmetry with respect to any\nplane containing a pair of semi-axes, the angle \u00121is equivalent to \u00121+k\u0019,k2Z. In\nconsequence, if m= 2k1;k12Z, type 00is equivalent to type 0 and type\u0019\n20is equivalent\nto type\u0019\n2. Likewise, for m= 2k2+ 1,k22Z, type 00is equivalent to type\u0019\n2and type\u0019\n20\nis equivalent to type 0. Then, for resonances of order m: 2,m2Z, we will take types 0\nand\u0019\n2as representatives. With this we have proved item 2.\nThe previous facts rely only on the symmetries and the periodicity of equation (23),\nincluding the discussion in [CC00]. Since equation (25) for the spin motion of E1, with\nE2spherical, has exactly the same properties, then, the proof above is also valid in such\ncase. \u0003\nProposition 5 let us characterize all the balanced spin-orbit resonances in the \frst\nhalf of an orbital revolution, additionally, the corresponding solutions have a certain\nsymmetry relation. In the generalization to spin-spin resonances we want to combine\ndi\u000berent spin-orbit resonances and we will use the boundary conditions in the same time\ninterval.\nRemark 6.Note that solutions of type\u0019\n2can be recovered with the conditions of type 0\nby considering negative values of \u00151. More precisely, solutions of type\u0019\n2of equation (23)\nwith\u00151=\u0015\u0003>0 are equivalent to solutions of (23) with \u00151=\u0000\u0015\u0003satisfying conditions\nof type 0. The same holds for (25).\nRemark 7.Proposition 5 gives a way to numerically search for balanced resonances.\nIndeed, eqs. (34) and (35) can be used to apply a Newton method, that is, to \fnd the\ninitial condition _\u00121(0) such that the conditions for either Type 0 or Type\u0019\n2are satis\fed.\nNext we introduce the following de\fnition, which deals with the spins of both objects.14 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nDe\fnition 8. We say that the ellipsoids E1,E2are in a standard spin-spin resonance3of\norder (m1:n1;m2:n2) withm1;m22Z;n1;n22Znf0g, if the ellipsoidEjis in amj:nj\nspin-orbit resonance. In such case, the resonant angles m1:n1m2:n2(t) = m1:n1\n1(t)\u0006 m2:n2\n2(t)\nare 2\u0019n-periodic functions, where nis the least common multiple of n1andn2.\nAn analogous de\fnition holds for resonances of the balanced type.\nDe\fnition 9. We say that the ellipsoids E1,E2are in a balanced spin-spin resonance\nof order (m1: 2;m2: 2) withm1;m22Z, if the ellipsoidEjis in amj: 2 spin-orbit\nresonance for j= 1;2.\nNote that a spin-spin resonance of order ( m1:n1;m2:n2) is also of order ( \u00141m1:\nn;\u0014 2m2:n), wheren=\u00141n1=\u00142n2is the least common multiple of n1andn2. Again,\nthe converse is not true in general. For example, the resonance (1 : 1 ;3 : 2) is of order\n(2 : 2;3 : 2), but not the opposite. However, a balanced resonance (2 : 2 ;3 : 2) is a\nspin-spin resonance of order (1 : 1 ;3 : 2).\nThe following proposition generalizes Proposition 5 to the spin-spin problem.\nProposition 10. The following statements hold for the spin-spin problem (24):\n(i)AnyR\u000b;\f1;\f2-symmetric orbit, with R\u000b;\f1;\f2de\fned in (30), associated to a (m1:\nn;m 2:n)spin-orbit resonance is equivalent to a solution that satis\fes the fol-\nlowing Dirichlet conditions:\n\u0012j(\u000b) =\fj; \u0012j(\u000b+n\u0019) =\fj+mj\u0019 ;\nwith\u000b2f0;\u0019gand\fj2f0;\u0019\n2g(eight combinations). Moreover, such solution\nsatis\fes the following symmetry property \u0012j(t) = 2\fj\u0000\u0012j(2\u000b\u0000t).\n(ii)There are four independent types of R\u000b;\f1;\f2-symmetric orbits representing a bal-\nanced (m1: 2;m2: 2)spin-spin resonance and are given by:\nType (\f1;\f2) :\u0012j(0) =\fj; \u0012j(\u0019) =\fj+mj\u0019\n2; (39)\nwith\fj2 f0;\u0019\n2g. Moreover, the corresponding symmetry relation is \u0012j(t) =\n2\fj\u0000\u0012j(\u0000t).\nProof. This proof is based on the fact that the same arguments used to prove Proposi-\ntion 5 can be generalized in a straightforward way for the spin-spin problem (24).\nNow we apply Lemma 2 to (26) with F(x) given by (29). The \fxed point set of each\nreversing symmetry R\u000b;\f1;\f2is\nFix(R\u000b;\f1;\f2) =f(t;\u00121;\u00122;p1;p2):t=\u000b;\u0012 1=\f1;\u00122=\f2g:\nThen, due to the periodicity of Win (20), the periodic orbits can be found at \u0012j(\u000b) =\fj,\nwhere we can take any combination between \u000b2f0;\u0019gand\fj2f0;\u0019\n2g. A similar\nmethod was used in [Gre79] for the standard map and was generalized for the spin-orbit\nproblem in [CC00] and for a standard map of two degrees of freedom in [CFL04].\nThe rest of the proof of item i follows analogously to the proof of item (1) of Proposi-\ntion 5 using the reversing symmetries.\n3We remark that this is a practical de\fnition because it is useful for the physical interpretation.\nHowever, we notice that there is a more general de\fnition given by the resonant combination n0t\u0000\nn1\u00121\u0000n2\u00122, withn0;n1;n2integers.THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 15\nThe proof of item ii needs more detail. In the same way as (38), we obtain easily that\nthe balanced resonances ( m1: 2;m2: 2) are given by\n\u0012j(\u000b) =\fj; \u0012j(\u000b\u0006\u0019) =\fj\u0006mj\u0019\n2: (40)\nWe see that (39) corresponds to (40) with \u000b= 0 and the positive sign. From the case\n\u000b=\u0019and the negative sign we have that\n\u0012j(0) =\fj\u0000mj\u0019\n2; \u0012j(\u0019) =\fj: (41)\nNote that if, for example, we take j= 1 andm1= 2k1, withk12Z, then, the conditions\n(39) and (41) are equivalent. On the other hand, if m1= 2k1+ 1, then, the conditions\n(39) with\f1= 0 and\u0019\n2are equivalent respectively to (41) with \f1=\u0019\n2and 0. The same\nis true forj= 2. Then, with (39) all the possibilities are covered. \u0003\nRemark 11.Results in Proposition 10 allow to apply a Newton method to compute\nresonances for the spin-spin problem just considering as unknowns _\u0012j(0) and correct\nthem by imposing the conditions in (39) or (41).\nFor circular orbits ( e= 0), each of the spin-orbit models (23) is a classical pendulum\nwhose only stable equilibrium point corresponds to a 1 : 1 resonance that is given by\n\u0012j(t) =t. Similarly, the spin-spin model (24) consists of two coupled penduli whose only\nstable solution is \u00121(t) =\u00122(t) =tthat is a (1 : 1 ;1 : 1) resonance. However, for e6= 0,\nf(t;e) does not coincide with tand more stable spin-orbit resonances may appear.\nIn order to study the spin-orbit and spin-spin resonances, it is useful to compute the\nexpansion of V2andV4up to some power of the eccentricity. This expansion is obtained\nsolving Kepler's equation (13) up to a \fnite order in the eccentricity, then inserting the\nsolutionu=u(t) in (14), (16), expand them in series of the eccentricity and \fnally\nexpanding the trigonometric terms appearing in (8).\nThis procedure leads to the expansions of V2andV4that, for simplicity, we give up to\nthe order 2 in the eccentricity in the Appendix C. In those expressions, the trigonometric\nterms with arguments mj:nj\nj(t) =mjt\u0000nj\u0012jare associated to mj:njspin-orbit reso-\nnances forEj, whereas the terms with argument m1:n1m2:n2(t) = (m1\u0006m2)t\u0000n1\u00121(t)\u0007n2\u00122(t)\ncorrespond to spin-spin resonances by combining spin-orbit resonances of orders m1:n1\nandm2:n2. For each order of the expansion in the eccentricity e, there are some reso-\nnances appearing. They are shown in Table 1, where we can recognize a hierarchy: the\nmost important spin-orbit resonance is 1 : 1, then we have 3 : 2, 1 : 2 and so on, because\nthey appear for low orders of the eccentricity. Resonances of further orders are relevant\nonly for large eccentricities.\nNote that the most relevant spin-orbit resonances are of order m: 2,m2Z. Ad-\nditionally, spin-spin resonances appearing at order e\u000binV4are obtained by combining\nspin-orbit resonances appearing at order e\u000b1ande\u000b2inV2such that\u000b1+\u000b2=\u000b.\nFinally, note that for V2, the coe\u000ecients associated to spin-orbit resonances are of order\none ind1;d2, whereas for V4, the spin-orbit coe\u000ecients are of order two in d1;d2;q1;q2,\nand the spin-spin coupling coe\u000ecients are of order two in d1;d2.\n3.2.KAM tori in the spin-spin problem. Now we deal with quasi-periodic solutions\nof the Keplerian version of the spin-spin model. We denote by\n!= (1;!1;!2) (42)16 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nspin-orbit resonances spin-spin resonances\norder V2 V4 V4\ne01 : 1 1 : 1 combine 1 : 1 with 1 : 1\ne13 : 2, 1 : 2 3 : 2, 1 : 2, 3 : 4, 5 : 4 combine 1 : 1 with 3 : 2 and 1 : 2\ne21 : 1, 2 : 1, 0 : 11 : 1, 2 : 1, 0 : 1,\n1 : 2, 3 : 2combine 1 : 1 with 1 : 1, 2 : 1 and 0 : 1\nall combinations between 3 : 2 and 1 : 2\nTable 1. Resonances appearing in the expansion of the potential V2+V4\nfor each order of the eccentricity.\nthe frequency vector with\n!1=p1\nC1; ! 2=p2\nC2:\nThe Hessian matrix associated to (20) has determinant di\u000berent from zero, whenever\np16= 0,p26= 0. This implies that (20) satis\fes the Kolmogorov non-degeneracy condition,\nwhich is a requirement for the applicability of KAM theorem [Kol54, Arn63, Mos62]. The\nother essential requirement in KAM theory is the assumption that the frequency satis\fes\na Diophantine inequality, namely there exist C > 0 and\u0018\u00152, such that\njk\u0001!j\u00001\u0014Cjkj\u0018(43)\nfork2Z3nf0g.\nWe remark that a possible choice of !satisfying (43) can be obtained as follows. Let\n\u000bbe an algebraic number of degree 3, namely the solution of a polynomial equation of\ndegree 3 with integer coe\u000ecients, not being the root of polynomial equations of lower\ndegree. Let us consider the vector != (1;!1;!2) obtained as\n0\nBBB@1\n!1\n!21\nCCCA=0\nBBB@1 0 0\nb1a11a12\nb2a21a221\nCCCA0\nBBB@1\n\u000b\n\u000b21\nCCCA; (44)\nwhere (b1;b2) and the matrix A\u0011(amn) have rational coe\u000ecients and det A6= 0. By\nnumber theory results (see, e.g., [CFL04]), a vector !as in (44) satis\fes (43) with\n\u0018= 2. Under smallness conditions of the parameters, say \u0015jin (22), KAM theory ensures\nthe existence of a quasi-periodic torus with Diophantine frequency. We remark that\nthe theory presented in [CCGdlL21c] for the spin-orbit problem (see also [CCGdlL21b,\nCCGdlL21a]) can be extended to provide explicit estimates for (20) and an explicit\nalgorithm to construct quasi-periodic solutions.\n4.Qualitative description of the spin models\nIn this section, we give a qualitative description of the phase space associated to\nthe spin-orbit problem (Section 4.1), the spin-spin problem with spherical companion\n(Section 4.2) and with non-spherical companion (Section 4.3).THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 17\n4.1.Spin-orbit problem ( Vper=V2).Since the system (23) corresponds to two uncou-\npled spin-orbit problems, then the Poincar\u0013 e map of the whole system can be understood\nas the direct product of the Poincar\u0013 e maps for each of the bodies. Consider for exam-\nple the dynamics of E1. Figure 3 is a typical Poincar\u0013 e map of the spin-orbit problem\nobtained using a Taylor integrator [JZ05] and using a similar approach as the one ex-\nplained in Appendix B, in this case with parameters ( e;\u00151) = (0:06;0:05); it represents\nsolutions at t= 2\u0019k,k2Z, in the plane ( \u00121;_\u00121) restricted to \u001212[\u0000\u0019;\u0019]. The main\nstable resonances are tagged with their corresponding order m:n. The Poincar\u0013 e map\nfor su\u000eciently small parameters ( e;\u00151) has the following features:\n1) The main stable spin-orbit resonances are represented by \fxed points in the plane\n(\u00121;_\u00121) surrounded by islands of invariant librational tori. High order resonances\nappear above low order resonances.\n2) It looks that the spin-orbit resonances of order m: 2\u00151 : 1 are balanced:\nsolutions of type 0 (34) are stable and those of type\u0019\n2(35) are unstable. On\nthe contrary, for the 1 : 2 resonance, type 0 is unstable and type\u0019\n2is stable.\nConcerning other resonances (33), for instance in the 3 : 4 resonance, types with\n\u000b= 0 are unstable and types with \u000b=\u0019are stable. Exactly the opposite occurs\nfor the case 5 : 4.\n3) It is possible to have stable secondary resonances, namely small resonances sur-\nrounding other resonances. This is clear for the 1 : 1 resonance. Beyond the\nlibrational islands associated to the resonance 1 : 1 there is a chaotic region that\nincludes the unstable resonances and that is larger for large parameters ( e;\u00151).\nThe chaotic region can appear for other resonances and is limited by rotational\ntori that also distinguish the domains of resonances of di\u000berent orders.\nFigure 3. Poincar\u0013 e map for the spin-orbit problem.\n4.2.Spin-spin problem ( Vper=V2+V4) with spherical companion. Let us now\nconsider the case when E2is a sphere. Then, \u00122(t) =\u00122(0) + _\u00122(0)tand the dynamics of\n\u00121is given by (25), that depends on the parameters ( e;\u00151;^q1;\u001b1). Here the parameters18 A. CELLETTI, J. GIMENO, AND M. MISQUERO\n\u001b1and ^q1perturb the previous framework of the spin-orbit problem, see Section 4.1.\nWe can see a comparison between both problems in Figure 4: we see that the only\nqualitative di\u000berence between both cases is that the Poincar\u0013 e map associated to equation\n(25) is slightly more chaotic. This minor di\u000berence is due to the fact that we take\n\u001b1= ^q1= 0:01, that are small parameters. As we will see in Section 6, the spin-spin\nmodel is a good approximation for the dynamics of two ellipsoids for \u001bjand ^q2up to\nthe order of magnitude of about 10\u00002, because larger values could lead to a collision, see\nSection 6.2.\nFigure 4. Poincar\u0013 e maps. Left: usual spin-orbit problem with\n(e;\u00151) = (0:06;0:05). Right: spin-orbit problem up to order 1 =r5with\n(e;\u00151;\u001b1;^q1) = (0:06;0:05;0:01;0:01).\n4.3.Spin-spin problem ( Vper=V2+V4) with non-spherical companion. Now we\ndeal with the general system (24). Let \t( t) = (\u00121(t);\u00122(t);_\u00121(t);_\u00122(t)) be a solution of\n(24) and its respective projections \t 1(t) = (\u00121(t);_\u00121(t)) and \t 2(t) = (\u00122(t);_\u00122(t)). From\nnow on we restrict \u0012j(t) to [\u0000\u0019;\u0019]. Let the Poincar\u0013 e map associated to such solution\nbe de\fned byP(\t(0)) = \t(2 \u0019), and its projections by Pj(\tj(0)) = \t j(2\u0019). It is\nnot possible to represent the iteration of the map Pin a single plot, because it is 4-\ndimensional, so we will represent the projections Pj. That is to say, for a solution \t( t),\nwe are interested in the behavior of the two families of points \t 1(2\u0019k) and \t 2(2\u0019k),\nk2Z, in a single 2-dimensional strip \u0005 = f(x;y)2R2:jxj\u0014\u0019g. We recognize the\nfollowing features:\n1) A solution in spin-spin resonance corresponds to a family of isolated recurrent\npoints ofP(namely, the successive points on the Poincar\u0013 e map), whose projec-\ntions are represented in \u0005 as a pair of families of recurrent points.\n2) If the spin-spin resonance is stable, then nearby solutions would librate around\nsuch points. While in the uncoupled system, librating solutions belong to 2-\ndimensional tori, here tori can be higher dimensional. As a result, the projected\npoints represented in \u0005 are distributed in two clouds of points surrounding each\nrecurrent point. A cloud of this kind covers an annulus-like region of a certain\nthickness that is usually thicker for stronger couplings. A similar behavior occursTHE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 19\nfor rotational solutions, whose corresponding clouds are distributed in strips of\ncertain thickness, see Figure 5.\n 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 1.6 1.7\n-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1dθ/dt(t)\nθ(t)/πσ=0.2 θ 1(0)=0 dθ 1/dt(0)=0.9 θ 2(0)=0 dθ 2/dt(0)=1.45\nbody 1\nbody 2\n 0.4 0.6 0.8 1 1.2 1.4 1.6\n-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1dθ/dt(t)\nθ(t)/πσ=0.25 θ 1(0)=0 dθ 1/dt(0)=0.5 θ 2(0)=0 dθ 2/dt(0)=1.51\nbody 1\nbody 2\nFigure 5. Left: Projections \t 1(2\u0019k) and \t 2(2\u0019k),k2Nof a solution\n\t(2\u0019k) librating around a stable (1 : 1 ;3 : 2) spin-spin resonance. Right:\nSolution for which one of the bodies librates around a stable spin-orbit\nresonance and the other one has a rotational behavior. The common pa-\nrameter values are e= 0:06,\u00151=\u00152= 0:05, and ^q1= ^q2= 0:001.\n3) We expect that, for small enough coupling parameters \u001bj, the spin-spin resonances\nare located close to the corresponding spin-orbit resonances for each ellipsoid, and\nwould keep the same stability as for the uncoupled problem. However, we have\nfound that, for larger \u001bj, the stability may change with respect to the uncoupled\nsystem, see Figure 6.\n4) The coupled system is 5-dimensional, then, invariant tori, if there exist, would\nnot con\fne solutions in determined regions (as in the uncoupled system), but\nArnold di\u000busion is expected to take place.\nA particular behavior occurs only when both bodies are identical ( C1=C2= 0:5,\n\u0015=\u00151=\u00152,\u001b=\u001b1=\u001b2and ^q= ^q1= ^q2), the so-called measure synchronization . This\nphenomenon was observed numerically in [HZ99] for an autonomous Hamiltonian system\nof two degrees of freedom (a pair of identical coupled oscillators): the system librates\naround a stable periodic solution in a very particular way described as follows for our\nsystem (see the phenomenon illustrated in Figure 7). Take a solution \t( t) librating\naround a stable spin-spin resonance of order ( m:n;m :n). Consider the two families\nof points \t 1(2\u0019k) and \t 2(2\u0019k) and their corresponding annulus-like region in the plane\n\u0005. There are two possibilities: either both clouds of points are distributed in separated\nannulus-like regions or both regions coincide. That is to say, either the overlap is empty\nor there is a complete overlap. Moreover, if we start with a solution with separated\nregions, we can obtain the complete overlap by increasing the coupling parameter \u001b\n(keeping the same \t(0)). The phenomenon takes place suddenly for a speci\fc \u001b=\u001b\u0003\nwhen the outer boundary of the inner ring touches the inner boundary of the outer ring.\nAt that moment, there is a concentration of density of points in the contact region.20 A. CELLETTI, J. GIMENO, AND M. MISQUERO\n 0.4 0.6 0.8 1 1.2 1.4 1.6\n-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1dθ/dt(t)\nθ(t)/πσ=0.15 θ 1(0)=1.571 dθ 1/dt(0)=0.48 θ 2(0)=0 dθ 2/dt(0)=1.48\nbody 1\nbody 2\n 0.4 0.6 0.8 1 1.2 1.4 1.6\n-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1dθ/dt(t)\nθ(t)/πσ=0.55 θ 1(0)=0 dθ 1/dt(0)=0.48 θ 2(0)=0 dθ 2/dt(0)=1.48\nbody 1\nbody 2\nFigure 6. Left: Projections \t 1(2\u0019k) and \t 2(2\u0019k),k2Zof a solution\n\t(t) close to an unstable (1 : 2 ;3 : 2) spin-spin resonance. Right: Rep-\nresentation of a solution with identical \t(0) for larger coupling, now the\nnearby spin-spin resonance has become stable. The common parameter\nvalues aree= 0:06,\u00151=\u00152= 0:05, and ^q1= ^q2= 0:001.\nThe phenomenon of measure synchronization disappears when the bodies are not\nequal. See in Figure 8 how the domains of both ellipsoids can overlap without merging\ninto a single ring.\n5.Linear stability of resonances\nIn this section we analyze the stability (in the linear approximation) of solutions asso-\nciated to the resonances in di\u000berent models, namely the spin-orbit problem (Section 5.1)\nand the spin-spin problem with spherical (Section 5.2) and non-spherical (Section 5.3)\ncompanion.\nIn all cases, we will only deal with balanced resonances (32), because they appear to\nbe simpler and more relevant (see Section 4) than resonances of the general type (31).\nActually, we can establish regions in the space of parameters where solutions associated\nto balanced resonances are unique or have some low multiplicity. In the case of the spin-\nspin problem, we restrict ourselves to regions of uniqueness. The study of linear stability\nof such periodic solutions in the space of parameters complements the understanding of\nthe dynamics that we presented in previous sections, especially Section 4.\n5.1.Spin-orbit problem ( Vper=V2).Consider the spin-orbit problem (23) with j= 1,\nthat is, the motion of the ellipsoid E1. Let\u00121=\u0012\u0003(t) be a solution in a balanced m: 2\nspin-orbit resonance, whose associated variational equation is\ny+\u00151\u0012a\nr(t;e)\u00133\ncos(2\u0012\u0003(t)\u00002f(t;e))y= 0; y2R: (45)\nFore6= 0, (45) is a linear equation with a 2 \u0019-periodic coe\u000ecient. Particularly, this is a\nHill's equation [MW79]. Assume that \b( t) is a matrix solution of (45) with \b( t0) =12,\nthe identity matrix 2 \u00022. The stability of (45) is determined by the structure of the\nmatrixM= \b(t0+ 2\u0019), called monodromy matrix. If jTr(M)j<2, we have ellipticTHE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 21\n 0.9 0.95 1 1.05 1.1\n-0.15 -0.1 -0.05 0 0.05 0.1 0.15dθ/dt(t)σ=0\nbody 1\nbody 2\n 0.9 0.95 1 1.05 1.1\n-0.15 -0.1 -0.05 0 0.05 0.1 0.15σ=(0.1,0.1)\nbody 1\nbody 2\n 0.9 0.95 1 1.05 1.1\n-0.15 -0.1 -0.05 0 0.05 0.1 0.15dθ/dt(t)\nθ(t)/πσ=0.268\nbody 1\nbody 2\n 0.9 0.95 1 1.05 1.1\n-0.15 -0.1 -0.05 0 0.05 0.1 0.15\nθ(t)/πσ=0.268884\nbody 1\nbody 2\nFigure 7. Projections \t 1(2\u0019k) and \t 2(2\u0019k),k2Zof a solution \t( t)\nclose to a stable (1 : 1 ;1 : 1) spin-spin resonance for di\u000berent values of\n\u001b1and\u001b2. Keeping the same parameters e= 0:06,\u00151=\u00152= 0:05,\n^q1= ^q2= 0:001,\u00121(0) =\u00122(0) = 0, _\u00121(0) = 0:92 and _\u00122(0) = 1:05. The\nexternal ring (body 1) keeps similar thickness, the internal ring changes\nfrom a thin one to another that occupies values close to (0,0) to thin one\nto \fnaly collapse with the exterior one.\nstability, whereas if jTr(M)j>2 we have hyperbolic instability. In the parabolic case,\nwhenjTr(M)j= 2, the system is stable if the Jordan canonical form of Mis12or\n\u000012, otherwise the system is unstable. Actually, if the system is parabolic unstable, the\ninstability of the linear system is linear in time, but hyperbolic instability is associated to\nan exponential divergence in time. In our case we want to distinguish regions of stability\nand instability in the ( e;\u00151)-plane for a given solution, which is continuous in ( e;\u00151).\nFrom properties of Hill's equations, regions of elliptic stability are separated from regions\nof hyperbolic instability by parabolic curves ( jTr(M)j= 2). These curves are made of\nunstable points, except if there are intersections of parabolic curves, because points of\nintersection become stable. This phenomenon is called coexistence , [MW79].22 A. CELLETTI, J. GIMENO, AND M. MISQUERO\n 0.9 0.92 0.94 0.96 0.98 1 1.02 1.04 1.06 1.08 1.1\n-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4dθ/dt(t)\nθ(t)/πbody 1\nbody 2\nFigure 8. Partial overlapping of the domains of \t 1(2\u0019k) and \t 2(2\u0019k),\nk2Zof a solution \t( t) close to a stable (1 : 1 ;1 : 1) spin-spin resonance in\nthe case of di\u000berent bodies: no measure synchronization. The parameter\nvalues aree= 0:06,\u00151= 0:009,\u00152= 0:05, ^q1= ^q2= 0:001,\u001b1=\u001b2= 0:3,\n\u00121(0) =\u00122(0) = 0, _\u00121(0) = 0:92, and _\u00122(0) = 1:064.\nFor a given point ( e;\u00151) and a given balanced m: 2 resonance, we want to know how\nmany solutions there are of each type (34) or (35), and their linear stability. First, recall\nRemark 6: solutions of type\u0019\n2satisfy conditions of type 0 for the equation (23) with j= 1,\ntaking\u0000\u00151instead of\u00151. Consequently, for each ( e;\u00151), withe2[0;1) and\u001512(\u00003;3),\nwe can obtain all the solutions corresponding to a balanced resonance by applying the\nshooting method: take a solution \u00121(t) with initial conditions4\u00121(\u0019) =m\u0019=2,_\u00121(\u0019) =\n\r2Rand let\rvary until the boundary condition \u00121(0) = 0 is reached. Finally, we obtain\nthe linear stability of the solution by computing \b( t) such that \b( \u0019) =12. Actually,\nfor this procedure, we can take any of the boundary conditions in eqs. (34) and (35), we\njust need one type to generate all solutions.\nThe results of this method for the main balanced spin-orbit resonances are shown in\nFigure 9. For these computations we used a Runge-Kutta Verner 8(9) integrator [Ver78],\ninstead of a Taylor integrator [JZ05]; the reason is that, for some parameter values, the\nsolutions are constant or polynomials and the Taylor method su\u000bers in choosing a good\nstep size. Thus, Figure 9 required around 3.5 days with 34 CPUs to be generated with\na discretization mesh size of 2000 \u00022000\u00022000 for (e;\u00151;\r).\nWe can recognize the following characteristics5:\n1) Each of the balanced resonances is represented in the ( e;\u00151;\r)-space by a contin-\nuous surface. In the case 1 : 1, the surface is made of two sheets connected only\nin one point ( e;\u00151) = (0;1).\n4We choose to take initial conditions at t=\u0019and nott= 0 because the values of _\u00121(0) producing\nspin-orbit resonances for large eand\u00151are too large to be represented in a 3-dimensional plot as in\nFigure 9.\n5The linear stability of the multiple solutions associated to the resonances 1 : 1 and 3 : 2 was already\nstudied in [ZOST66, Bel66, BL75], but we include them here in order to have a more complete view.THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 23\nFigure 9. Diagrams of linear stability for the balanced resonances of\norder 1 : 2 (left), 1 : 1 (middle left), 3 : 2 (middle right) and 2 : 1 (right).\nBlue-instability and yellow-stability. The upper diagrams are projections\nof the lower diagrams in the ( e;\u00151)-plane with some transparency in order\nto identify regions with multiplicity of solutions.\n2) The region of uniqueness in the ( e;\u00151)-plane is quite large. The multiplicity\nis generated by bifurcation of solutions: the surface folds generating multiple\nsolutions (from one to three, as far as we see). Particularly, the bifurcation in the\ncase 1 : 1 occurs at ( e;\u00151) = (0;1) producing a secondary sheet behind the main\none. In general, the multiplicity takes place for some regions with j\u00151j>1. The\nresonances of order m: 2 withm > 2, have two characteristic folds, one with a\nV-like shape in solutions of type 0 for \u00151>1, and another small one in solutions\nof type\u0019\n2for\u00151\u0018\u00002 and very large eccentricities.\n3) Instability is predominant in the diagrams, especially in solutions of type 0 for\nthe resonance 1:2 and of type\u0019\n2for the rest of the resonances. We see that the\nmain regions of linear stability are continuation of stable solutions from e= 0,\nmuch of which are close to small j\u00151j. Except for the resonance 1:2, the stability\nregion for large eccentricities ( e>0:6) of the other resonances has a similar shape,\ncharacterized by a bifurcation with an interchange of stability. It is interesting\nto note that the folds producing the multiplicity are associated to some stable\nregions with peculiar shapes. In the resonance 1 : 1 we \fnd two unstable regions\nbifurcating from the exact solution \u00121(t) =tfore= 0: one at \u00151= 1=4 = 0:25\n(main sheet) and the other one at \u00151= 9=4 = 2:25 (secondary sheet).\nNow let us consider both bodies. Since the system (23) is uncoupled, then, the multi-\nplicity and stability associated to a spin-spin resonance is given by each of the separated\nproblems. For example, take the (1 : 1 ;3 : 2) balanced spin-spin resonance of type (0 ;\u0019\n2)24 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nfor (e;\u00151;\u00152) = (0:3;0:1;0:5), that is, the red points shown in Figure 9. It has associated\na unique solution and it is unstable because it is so for \u00122.\n5.2.Spin-spin problem ( Vper=V2+V4) with spherical companion. In this case\nwe know thatE2is in uniform rotation \u00122(t) =\u00122(0) + _\u00122(0)t, while the dynamics of\n\u00121is given by (25), that depends on ( e;\u00151;^q1;\u001b1). For this problem, we can proceed in\nthe same way as for the spin-orbit problem of Section 5.1. On one hand, the variational\nequation associated to a solution in a spin-orbit resonance is a Hill's equation like (45),\nso the linear stability of the solution is characterized by the corresponding monodromy\nmatrix. On the other hand, we can \fnd all the solutions associated to a balanced spin-\norbit resonance using the shooting method for only one type of boundary conditions by\nincluding negative values of \u00151, see Remark 6.\nComparing Figures 9 and 10 we can see, for example, how the balanced 3 : 2 resonance\nis perturbed when we turn on the parameters (^ q1;\u001b1):\nFigure 10. Diagrams of linear stability for the 3 : 2 balanced resonance of\n(25) for di\u000berent values of the indicated parameters (^ q1;\u001b1). Blue denotes\ninstability and yellow denotes stability. Each of the columns have required\naround 15h using 24 CPUs and a mesh size of 512 \u0002512\u0002512.\n1) The e\u000bect of the new parameters is remarkable for large eandj\u00151j. Especially\nwhen ^q1and\u001b1are large.\n2) At very large eccentricities, the surface has a complicated structure resulting in\nmultiple solutions. The e\u000bect of ^ qis mainly to alter the stability for large e:\nsolutions of type\u0019\n2become always unstable, while stable regions of solutions of\ntype 0 are more concentrated. The growth of ^ qalso increases the multiplicity of\nsolutions of type 0 and very large e. On the other hand, increasing \u001b1has a more\ndramatic e\u000bect on the complexity of the surface and also modi\fes the stability for\nlargee. Actually, for some values of \u001b1, the existing V-shaped fold connects with\nthe complex structure of large eccentricities in the upper part of the diagram.THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 25\n5.3.Spin-spin problem ( Vper=V2+V4) with non-spherical companion. We study\nthe linear stability of the resonances of the full coupled spin-spin model in (24). In this\ncase linear stability is determined by a more general theory and we will restrict ourselves\nto zones in the space of parameters where there is uniqueness.\nSince we have two degrees of freedom, the \frst variation at a particular resonance is\nnot a Hill's equation anymore, but a coupled system of second order. A Hill's equation is\na particular case of linear Hamiltonian system with periodic coe\u000ecients, hereafter LPH\nsystems.\nIf we de\fne z= (\u00121;\u00122;p1;p2)T, wherepj=Cj_\u0012j, the spin-spin problem in the Hamil-\ntonian form (21) can be written as\n_z=J2@zHK(t;z); (46)\nwhereJlis the square matrix of order 2 lgiven by\nJl=0\n@01l\n\u00001l01\nA; l= 1;2;::: (47)\nwith 1lthe unit matrix of order l. The non-autonomous Hamiltonian HK(t;z) is given\nin (19) for Vper=V2+V4. Letz=z\u0003(t) be a solution of (46) that is in a balanced\nspin-spin resonance of type ( m1: 2;m2: 2). Then, the \frst variation at such solution\nhas the form\n_y=J2@z;zHK(t;z\u0003(t))y; y2R4; (48)\nwhere@z;zHKdenotes the Hessian matrix of the Hamiltonian HKin the 4-dimensional\nvariablez. The system (48) is an LPH system of period 2 \u0019. Assume that \b( t) is a\nmatrix solution of (48) with \b( t0) =14. The stability of (48) is determined by the\nFloquet multipliers, that are the eigenvalues of the monodromy matrix \b( t0+ 2\u0019).\nAn LPH system has particular stability properties, see the general theory in [YS75,\nEke90] and an application to the double synchronous spin-spin resonance in [Mis21].\nFor example, assume that '2Cis a Floquet multiplier of an LPH system. Then, its\ninverse'\u00001, its complex conjugates \u0016 'and \u0016'\u00001are also multipliers and have the same\nmultiplicity as '. That is, the Floquet multipliers have a symmetric distribution with\nrespect to the real line and the unit circle of the complex plane. In consequence, a\nnecessary condition for stability is that all multipliers have modulus 1. Moreover, if all\nmultipliers have modulus 1 and they are all di\u000berent, then the system is stable. When all\nthe multipliers have modulus 1 and some of them coincide, the situation is not trivial and\nthe stability depends on further algebraic properties of the multipliers (Krein's theory,\n[Kre50]). Then, unlike for Hill's equations, here we do not have a quantity like the trace\nof the monodromy matrix in order to characterize the boundary of stability/instability\nregions. Instead, we will use the following de\fnition.\nDe\fnition 12. We will say that an LPH system is hyperbolic unstable if\nmax\nkj'kj>1;\nwhere'k,k= 1;2;:::, are all the Floquet multipliers of the system.\nAssume that the solution z\u0003(t) is continuous in some domain of the parameters of the\nmodel (e;C1;\u00151;\u00152;\u001b1;^q1;^q2). The equation max\nk=1;:::;4j'kj= 1 +\", with\" >0, de\fnes a26 A. CELLETTI, J. GIMENO, AND M. MISQUERO\n1-parametric family of hyperbolic unstable manifolds in the space of parameters; then,\nthe boundary of hyperbolic instability will be found if we take the limit of such manifolds\nas\"!0.\nOn the other hand, it is possible to \fnd all the solutions of a given type of a balanced\nspin-spin resonance using the shooting method as in the previous section. However, since\nthe phase space and the space of parameters have large dimensions, our approach will be\nto obtain the solutions by continuation. Note that the solutions of (24) for \u0015j= 0 and\nanye2[0;1) are exactly given by \u0012j(t) =\u0012j(0) + _\u0012j(0)t. Then, the unique solution of\ntype (\f1;\f2) of a balanced spin-spin resonance ( m1: 2;m2: 2) is just \u0012j(t) =\fj+mj\n2t.\nSuch solution can be continued for j\u0015jj>0. For small enough j\u0015jj, the systems in\neqs. (23) to (25) can be regarded as di\u000berent perturbations of the system \u0012j= 0, then,\nFigures 9 and 10 give us a quite clear idea of a region of uniqueness of a given type\nassociated to a balanced spin-spin resonance of (24). Moreover, such solution can be\nfound by continuation in the space of parameters.\nFigure 11. Regions of (hyperbolic) linear instability of solutions asso-\nciated to di\u000berent spin-spin resonances of the problem (24) (case: equal\nbodies) for di\u000berent values of the parameters (^ q;\u001b). Blue denotes instabil-\nity and yellow denotes stability. Each plot has required around 10 minutes\nusing 15 CPUs and a mesh size 750 \u0002750.THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 27\nCase of equal bodies E1=E2.In this case, the equation (24) depends on four parameters\n(e;\u0015;^q;\u001b), where\u0015=\u00151=\u00152, ^q= ^q1= ^q2and\u001b=\u001b1=\u001b2. From Figure 10, it looks\nthat a good choice for doing the continuation is in the range 0 \u0014e.0:8, 0\u0014\u0015.1 and\n0\u0014^q;\u001b.0:01. In this range, the linear stability of the solution of type (0 ;0) associated\nto the resonance of order (1 : 1 ;1 : 1) was investigated in [Mis21]. Figure 11 shows\nthe stability diagrams of the resonance (1 : 1 ;3 : 2) of type (0 ;0) and the resonance\n(1 : 2;3 : 2) of type (\u0019\n2;0) and (0;0) for di\u000berent (^ q;\u001b). Let us point out some properties:\n1) Note that a plot with (^ q;\u001b) = (0;0) is just the superposition of diagrams in\nFigure 9, while a plot with (^ q;\u001b) = (0:01;0) is the superposition of diagrams\nsimilar to those in Figure 10. Then, the e\u000bect of the coupling can be seen in the\nplots with\u001b6= 0.\n2) The e\u000bect of the coupling is di\u000berent in each case: for the resonance (1 : 2 ;3 : 2)\nof type (0;0), we only see an additional thin unstable region for e < 0:1 and\n0:25<\u0015< 1. For the resonance (1 : 2 ;3 : 2) of type (\u0019\n2;0), we see that the region\nwith small ebecomes unstable, whereas for large ethe stability is somewhat\naltered. Finally, without coupling, the resonance (1 : 2 ;3 : 2) of type (0 ;0) is\nunstable for almost all the points, but the coupling introduces the stability for\nsmalle. Note that this in agreement with Figure 6.\n6.Comparison between the full and the Keplerian models\nIn this section, we provide some results on the comparison between the full and Ke-\nplerian problems (Section 6.1), and we give some numerical results on the interaction\nbetween the spin and orbital motions (Section 6.2).\n6.1.Hamiltonian approach. It will be useful to write the dynamical equations of the\nHamiltonian of the full model (4) in the compact form\n_z=J4@zH(z); (49)\nwherez= (r;f;\u0012 1;\u00122;pr;pf;p1;p2)TandJ4is de\fned by (47).\nLet us now formulate the Keplerian models (spin-orbit and spin-spin, see Section 2.3)\nas perturbations of the full model, so we can compare both families of models. Consider\na function\u0010(t) that satis\fes the equation\n_\u0010(t) =J4@zH(\u0010(t))\u0000h(\u0010(t)); (50)\nwhere the vector function\nh(z) = (0;0;0;0;\u0000@rVper(z);\u0000@fVper(z);0;0)T\nis responsible of subtracting the perturbative part of the potential (see equation (7))\nonly in the equations for _ prand _pf. Thus, equation (50) represents the Keplerian models\nincluding the orbital part: the spin-orbit model corresponds to (50) with Vper=V2and\nthe spin-spin model with Vper=V2+V4. The corresponding equations of motion are (11)\nand (12), so we can write the solution in the form\n\u0010(t) = (r(t;a;e);f(t;e);\u00121(t);\u00122(t);pr(t;a;e);pf(a;e);p1(t);p2(t));\nwhereaandeare the semimajor axis and the eccentricity of the Keplerian orbit, see\n(17). Now it is clear that the Keplerian model (50) is not Hamiltonian, even though we\ncan split it into one autonomous Hamiltonian system (orbital part with V=V0) and28 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nanother non-autonomous one (spin part with V=V0+Vper). Moreover, unlike for the\nfull model, in a Keplerian model neither the total angular momentum Pf=pf+p1+p2\nis conserved6, since we can compute that\n_Pf(t) =\u00002X\nj=1@\u0012jVper(r(t;a;e);f(t;e);\u00121(t);\u00122(t)):\nWe want to know if \u0010(t) is a good approximation for a solution of the full problem (49)\nwith the same initial conditions. Consider the solution z=z(t) =\u0010(t)\u0000\u000ez(t) of (49)\nsuch thatz(0) =\u0010(0). Then, expanding the equation (49) up to \frst order in \u000ez, we\nobtain that the function \u000ez=\u000ez(t) satis\fes the equation\nd\ndt\u000ez=J4@2\nz;zH(\u0010(t))\u000ez\u0000h(\u0010(t)) (51)\nwith initial condition \u000ez(0) = 0. In equation (51), @2\nz;zH(z) is the Hessian matrix associ-\nated to the Hamiltonian in (4). If the system (51) is stable, then the norm jj\u0010(t)\u0000z(t)jj\nis bounded. Additionally, it is easy to see that the system (51) is Lyapunov stable if and\nonly if the trivial solution of the homogeneous part\n_y=J4@2\nz;zH(\u0010(t))y; y2R8; (52)\nis Lyapunov stable. A general form of \u0010(t) is unknown because, although the orbital\npart is given by the classical Kepler problem, the spin part is given by a non-autonomous\nperiodic system of nonlinear equations. However, \u0010(t) is a periodic solution at spin-spin\nresonances, with period 2 \u0019for balanced resonances. In such cases, equation (52) is an\nLPH system, some of whose properties were mentioned in Section 5.3.\nAt this point, we wonder if it is possible for the periodic system (52) to be stable. Let us\npoint out an argument supporting a negative answer, even for stable spin-spin resonances.\nIt is known that the periodic solutions of the planar Kepler problem are not linearly\nstable. This can be easily checked using Poincar\u0013 e variables (action-angle variables): we\nobtain a positive eigenvalue for a linear system of constant coe\u000ecients. See further\nrelated discussions in [BO16, Sch72]. We can see that 1 is the only associated Floquet\nmultiplier and has multiplicity four; then, the instability of the periodic solutions of the\nKepler problem is not hyperbolic, according to De\fnition 12, but rather of parabolic\nkind. From this discussion, we expect that \u0010(t) andz(t) are divergent functions, but we\nwant to know if such divergence is exponential in time, that is, when (52) is hyperbolic\nunstable.\nSuppose that the function \u0010(t) is continuous on a domain of the space of parameters\n(a;e;C1;\u00151;\u00152;\u001b1;^q1;^q2). Note that we add ato the parameters of the spin-spin model\nbecause it varies with the orbital initial conditions. On the other hand, the Hamil-\ntonianHdepends on the parameters of the full model, we take the independent set\n(\u0016;C1;d1;d2;q1;q2), where the value of \u0016informs us about the disparity in the masses of\nthe bodies because \u0016=M1M2=M1(1\u0000M1). We can obtain ( \u0016;C1;d1;d2;q1;q2) from\n(a;e;C1;\u00151;\u00152;\u001b1;^q1;^q2) as follows: First, from (22), we see that, assuming \u001b1>0,\n\u0016=C1\n3\u001b1a2; (53)\n6Instead, the Keplerian assumption results in the conservation of the orbital angular momentum\npf(t) =\u0016r(t;a;e)2_f(t;e) =\u0016a2p\n1\u0000e2.THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 29\nfrom this value we can obtain M1because\u0016=M1(1\u0000M1) withM12(0;1). This is\nenough to write M2= 1\u0000M1,C2= 1\u0000C 1and, from the de\fnitions (22), it holds that\nd1=\u00151C1\n3M2; d 2=\u00152C2\n3M1; q 1= ^q1M1a2; q 2= ^q2M2a2:\nFrom these relations, it is obvious that, in order to compare the full and the Keplerian\nmodels, the spin-orbit versions are not enough, and we need to take Vper=V2+V4, say,\nthe spin-spin models.\nAdditionally, the initial conditions associated to \u0010(t) are given by (18), for the orbital\npart, and the values \u0012j(0);_\u0012j(0) are such that the spin part satis\fes the boundary con-\nditions for the spin-spin problem in a spin-spin resonance of a certain type as in (39).\nRecall also that, in our units, the Keplerian orbital period is T= 2\u0019andG=a3.\nAs in Section 5.3, we can \fnd the region (in the parameters space) of hyperbolic\ninstability and its boundary, that is given by the limit as \"!0 of the manifolds that\nsatisfy max\nk=1;:::;8j'kj= 1 +\", where'kare the Floquet multipliers of (52). It is important\nthat we focus on a region of the parameters where the corresponding solution associated\nto a spin-spin resonance is linearly stable, so we can see what is the e\u000bect of putting\nthe orbital and the spin parts altogether. Actually, not even the spin part of (52) is\ncompletely equivalent to (48), because all the derivatives of Vper(r;f;\u0012 1;\u00122), evaluated at\n\u0010(t), appear in (52) for both orbital and spin variables.\nFigure 12. Regions of (hyperbolic) linear instability of (1 : 2 ;3 : 2) of\ntype (\u0019=2;0) of the equation (52) for equal bodies and di\u000berent values\nof the parameters. Blue denotes instability and yellow denotes stability.\nEach plot has required around 10 minutes using 15 CPUs and a mesh size\nof 750\u0002750.\nFigure 12 shows the regions of hyperbolic instability of (52) in the case of equal bodies\nfor a particular spin-spin resonance with di\u000berent values of the parameters. We observe\nthe following:30 A. CELLETTI, J. GIMENO, AND M. MISQUERO\n(1) We can compare Figure 12 with Figure 11 for similar values of ^ qand\u001b. In the\ncase of equal bodies ( \u0016= 0:25 andC1= 0:5), the value of ais determined by\n(53): smaller \u001balso implies further bodies. Then, the plots with \u001b= 0:0001 in\nFigure 12 and be compared approximately with those with \u001b= 0 in Figure 11.\n(2) From the comparison, we see that the plots in Figure 12 with Figure 11 with\n^q= 0 are similar for small eccentricities, but, for large e, the unstable regions\nof plots in Figure 12 are larger. On the other hand, the plots in Figure 12 with\n^q= 0:01 include unstable stripes invading the stable regions for small e(they\ngrow with\u001b), whereas for large ethe diagrams become totally unstable.\n(3) As we increase \u001b(closer bodies), the plots become more and more unstable.\nWe conclude that the spin-spin model should be a reliable simpli\fcation of the full\nmodel in the regions of the parameters with small e, largeaand close to regions with\nstable spin-spin resonances. We remark also that instability of spin-spin resonances not\nnecessarily implies hyperbolic instability of (48): there are some regions for which the\nFloquet multipliers of the spin-spin resonance are not too far from the unit circle, so\nthat when we consider the corresponding system (48), all its Floquet multipliers belong\nto the unit circle. It is also remarkable that, although the comparison makes sense only\nbetween the full and Keplerian spin-spin models, we obtain almost identical plots as in\nFigure 12 independently if we use Vper=V2orVper=V2+V4in the Hamiltonian in (52).\n6.2.Quantitative numerical approach. In this section, we want to investigate, from\na numerical point of view, two questions that the Hamiltonian approach left unanswered:\n(Q1) Can we quantify the in\ruence of the spin motion on the orbital one? (Q2) Is it\npossible that the bodies end up colliding, even though (52) is not hyperbolic unstable?\nIn Section 6.1, we saw that the solution of a Keplerian model \u0010(t) corresponding to a\nspin-spin resonance should diverge from a solution of the full model z(t) with identical\ninitial conditions. In the region of parameters of hyperbolic instability the divergence\nshould be exponential; however, in the rest of the parametric space, we want to quantify\nhow di\u000berent both motions are. So, let us take the point ( a;e;C1;\u00151;\u00152;\u001b1;^q1;^q2) and\ncompute the functions \u0010(t) andz(t) in order to compare them.\nThe orbital motion of \u0010(t) is characterized unambiguously by the set of Keplerian\nelements (a;e;! ), where!= 0 is the argument of the periapsis, together with t, the\nmean anomaly. On the other hand, let the orbital part of the solution z(t) of the full\nmodel be given by ( rF(t);fF(t)). Then, we can transform the orbital position rF(t) =\nrF(t) exp(ifF(t)) to the osculating Keplerian elements ( aF(t);eF(t);!F(t)) of the two-\nbody problem using the following expressions. From the geometrical identity j_rFj2=\nG\u0010\n2\nrF\u00001\naF\u0011\n, we obtain\naF(t) = \n2\nrF\u0000_r2\nF+_f2\nFr2\nF\nG!\u00001\n:\nNow let us de\fne the orbital angular momentum per unit mass hF=rF^_rF, where^\nis the vector product, and the eccentricity vector\neF(t) =_rF^hF\nG\u0000rF\nrF;THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 31\nwhose modulus is given by\neF(t) =s\n1\u0000r4\nF_f2\nF\nGaF:\nWhenevereF6= 0, we can de\fne !F2[\u0000\u0019;\u0019) as the polar angle of eF. The mean\nanomaly associated to the full model can be de\fned too, but we will not use it in our\nstudy. The balanced spin-spin resonance of order ( m1: 2;m2: 2) in the function \u0010(t) is\ncharacterized by the modi\fed resonant angles mj:2\nj(t) =mjf(t;e)\u00002\u0012j(t). Seemingly,\nthe spin part of the solution z(t) of the full model is given by ( \u00121;F(t);\u00122;F(t)), so, in\norder to compare with \u0010(t), let us de\fne mj:2\nj;F(t) =mjfF(t)\u00002\u0012j;F(t). Now de\fne the\nfunctions\n\u000ea(t) =aF(t)\u0000a\na; \u000ee(t) =eF(t)\u0000e; \u000eres;j(t) = mj:2\nj(t)\u0000 mj:2\nj;F(t); (54)\nwhere\u000eais the relative deviation in semi-major axis, while \u000eeand\u000eres;jare the absolute\ndeviations in eccentricity and resonant angles, respectively.\nFinally, recall from (1) that we have an expression for ajin terms of the parameters of\nthe model. Then, for our purpose, we will say that there is a collision if rF(t)\u0014a1+a2.\nNow we are in a position to compare the solutions of the full model with respect with\nthose of the Keplerian models.\nIn Figure 13 we see, for di\u000berent values of the parameters and the case of identical\nbodies, the comparison between z(t) and\u0010(t) in a resonance (1 : 1 ;3 : 2) of type (0 ;0).\nParticularly, we see the evolution of the \u000e-functions in (54) and some Kepler elements of\nz(t) in 100 revolutions. In addition, Table 2 informs us about the corresponding Floquet\nmultipliers of both z(t) and\u0010(t). We summarize the following observations:\n(1) We chose two cases for the comparison. In the \frst case, e= 0 and\u001b= 10\u00003, so\nthe bodies are relatively close to each other ( a\u001939) in circular orbits, whereas in\nthe second case, e= 0:1 and\u001b= 10\u00007, so the bodies are much further ( a\u00193900)\nin moderately elliptic orbits. For both cases, we took values of parameters whose\ncorresponding resonance is not hyperbolic unstable, say, \u0015= 0:05 and ^q= 0:01.\n(2) From Figure 13, we see that the \frst case has a more regular behavior than the\nsecond one. In the \frst case, the orbit of the full model oscillates regularly very\nclose to the circular orbit of the Keplerian model, with a precession that increases\nuniformly in average. Actually, in a shorter time scale, the eccentricity vector\ndescribes a circles passing through the origin. On the other hand, the orbit of\nthe second case is greatly irregular, with variations of size up to \u001825%. Even its\nprecession changes direction at some moment and the eccentricity vector swings\nchaotically in short time scales.\n(3) The variation of the angles is quite regular in both cases. Except for the reso-\nnant angle of the resonance 1:1 of the \frst case, which oscillates regularly with\nsmall amplitude, the rest of the resonant angles increase/decrease more or less\nconstantly, with larger variations in the second case.\n(4) The Floquet multipliers in Table 2 gives us more information about both cases.\nWe remark that we display the multipliers of \u0010(t) corresponding to spin and orbit\nseparately, say, the \frst two rows are the four multipliers of the spin-spin model\n(24), whereas the third one shows the four coincident multipliers of the Kepler\nproblem. We con\frm that the spin of \u0010(t) is elliptic stable in the two cases.32 A. CELLETTI, J. GIMENO, AND M. MISQUERO\n-0.0004-0.0002 0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012 0.0014 0.0016\n 0 20 40 60 80 100δa(u)\nu/(2π) 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035\n 0 20 40 60 80 100δe(u)\nu/(2π)\n 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035\n 0 1 2 3 4 5 6 7 8eF(u)\nu/(2π)-1-0.8-0.6-0.4-0.2 0 0.2 0.4 0.6 0.8 1\n 0 20 40 60 80 100ω(u)/(2π)\nu/(2π) 0 0.01 0.02 0.03eF(ω)\n-0.4-0.3-0.2-0.1 0 0.1 0.2 0.3 0.4\n 0 20 40 60 80 100δres,1(u)\nu/(2π)-90-80-70-60-50-40-30-20-10 0\n 0 20 40 60 80 100δres,2(u)\nu/(2π)-15-10-5 0 5 10 15 20 25\n 0 20 40 60 80 100\nu/(2π)fF(u) - f(u)\nθ1,F(u) - θ 1(u)\nθ2,F(u) - θ 2(u)\n-0.3-0.25-0.2-0.15-0.1-0.05 0 0.05 0.1 0.15 0.2 0.25\n 0 20 40 60 80 100δa(u)\nu/(2π)-0.03-0.02-0.01 0 0.01 0.02 0.03\n 0 20 40 60 80 100δe(u)\nu/(2π)\n 0.075 0.08 0.085 0.09 0.095 0.1 0.105 0.11\n 0 1 2 3 4 5 6 7 8eF(u)\nu/(2π)-0.7-0.6-0.5-0.4-0.3-0.2-0.1 0 0.1 0.2 0.3\n 0 20 40 60 80 100ω(u)/(2π)\nu/(2π) 0 0.05 0.1eF(ω)\n-100 0 100 200 300 400 500 600 700 800 900\n 0 20 40 60 80 100δres,1(u)\nu/(2π)-900-800-700-600-500-400-300-200-100 0 100\n 0 20 40 60 80 100δres,2(u)\nu/(2π)-400-300-200-100 0 100 200 300 400 500\n 0 20 40 60 80 100\nu/(2π)fF(u) - f(u)\nθ1,F(u) - θ 1(u)\nθ2,F(u) - θ 2(u)\nFigure 13. Comparison between the full and the Keplerian solutions in\na resonance (1 : 1 ;3 : 2) of type (0 ;0) for di\u000berent parameter values in the\ncase of equal bodies. Top plots: e= 0,\u0015= 0:05, ^q= 0:01,\u001b= 10\u00003.\nBottom plots: e= 0:1,\u0015= 0:05, ^q= 0,\u001b= 10\u00007.THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 33\nOn the other hand, the multipliers of z(t) are computed with (49). We \fnd out\nthat, although the \frst case is more regular, the solution is actually hyperbolic\nunstable, whereas the opposite occurs for the second case: it is irregular but not\nhyperbolic. This is consistent with Figure 12.\ne= 0;\u0015= 0:05;^q= 0:01;\u001b= 10\u00003e= 0:1;\u0015= 0:05;^q= 0:01;\u001b= 10\u00007\nArgument Modulus Argument Modulus\n\u0010(t)\u00061:42 1\u00061:39 1\n\u00067:36\u000110\u000081\u00068:34\u000110\u000011\n0 1 (quadruple) 0 1 (quadruple)\nz(t)\u00061:41 1\u00061:39 1\n\u00067:64\u000110\u000091\u00068:34\u000110\u000011\n\u00061:65\u000110\u000011:09\u00063:77\u000110\u000041\n\u00061:65\u000110\u000010:915 0 1 (double)\nTable 2. Modulus and argument of the Floquet multipliers of the func-\ntionsz(t) and\u0010(t) compared in Figure 13.\nWe conclude that, even when there is hyperbolic instability, the solution of full system\nremains quite close to the solution of the Keplerian system with circular orbit. However,\neven with no hyperbolic instability, for e6= 0, the behavior of both solutions is remarkably\ndi\u000berent. This fact should be attenuated with very small eccentricities.\nWe can make a step further in the comparison when e= 0. Figure 14 shows a\nquantitative comparison in the orbits of the full and Keplerian models. We vary the\nparameters \u001band\u0015keeping ^q= 0 constant. For large enough \u001b(smalla), there is a\nregion of parameters leading to collision. The rest of the diagrams show di\u000berent orders\nof magnitude of \u000e-functions corresponding to aande. We check that, for very small \u0015,\nthe orbit of the full model is very close to the Keplerian one.\n7.Conclusions\nThis research is, primarily, a careful numerical study of the spin-spin model, presented\nas such in [Mis21]. For this purpose, we have considered several parallel models describing\nthe planar gravitational dynamics of two ellipsoids under the Assumptions 1 to 3. We\ndistinguish between full and Keplerian models, and also, between spin-orbit and spin-spin\nmodels. The known dynamics of the Keplerian spin-orbit problem was used as reference\nto develop our results.\nIn \frst place, we realize that the symmetries of the ellipsoids lead to certain symmetries\nin the equations of the Keplerian models, and so, to symmetric periodic solutions that\nstructure the overall dynamics. We complete this general view by providing conditions\nfor existence of quasi-periodic solutions. Between the periodic solutions, we focus on\na special class that comprises the simplest solutions from the physical interpretation,34 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nFigure 14. Contour plots of max flog10j\u000ee(u)j:u2[0;4\u0019]gon the left\nand maxflog10j\u000ea(u)j:u2[0;4\u0019]gon the right for the resonance (1 : 1 ;3 :\n2) of type (0 ;0) and parameter values e= 0 and ^q= 0 for equal bodies.\nThe black region denotes the collision during the integration of the full\nspin-spin model, i.e. when rF(u)\u0014a1+a2with ajde\fned in (1). Using\n80 CPUs the plots required around 1 hour with a mesh of 10242and 2048\nstopping points in [0 ;4\u0019].\nthe balanced spin-orbit and spin-spin resonances. We study the high-dimensional phase\nspace of the spin-spin problem by means of projections of Poincar\u0013 e maps, taking those\nof the spin-orbit problem as reference.\nFirst, we see that when one body is spherical, the dynamics of the spin-orbit problem\nis reproduced with small variations. For two ellipsoids, the projections of the Poincar\u0013 e\nmaps show structures similar to those of the spin-orbit ones. There is a particular\nbehavior when both bodies are identical, the so-called measure synchronization.\nAfter that, we study existence, multiplicity and linear stability of the solutions in bal-\nanced resonances as we vary the parameters of the problem. We focus on the qualitative\nchanges in the stability diagrams produced by the variation of each parameter. We use\nthe case of identical bodies to illustrate our observations most of the time.\nIn the last part of the paper, we compare the solutions of the full and the Keplerian\nmodels by means of rewriting the equations in a Hamiltonian-like form. We produce\nstability diagrams similar to the previous ones to show such a comparison. Representing\ndi\u000berent aspects with similar diagrams allows one a much more global understanding\nof a problem with so many variables and parameters. We end up by comparing the\nevolution of particular solutions for certain parameters. The orbit of the full problem is\ncharacterized by Kepler's elements to facilitate the comparison. From the parameters,\nwe observe that the comparison only makes sense for the spin-spin models. Here we \fnd\nout that, beside the case of circular orbits, for which full and Keplerian models show\na close behavior, for eccentric orbits, the trajectories are very dissimilar. We \fnish the\ncomparison in the case of circular orbits by showing that the space of parameters isTHE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 35\ndivided into regions leading to collision or regions with orbits di\u000bering speci\fc orders of\nmagnitude.\nThe topics and results presented in this work are just a taster of some features of the\nrich spin-spin dynamics. We also include a comparison with the full models, which is not\nwidely studied in the literature; however, this is crucial, because it shows us the limits\nof validity of the models for applications. We believe that here is still a lot to explore\nin this model, which can be used to investigate problems like Arnold's di\u000busion or the\nexistence of normally hyperbolic manifolds.\nAppendix A.Coefficients of the expansion of the potential\nThe full expansion of the potential energy of the Full Two-Body Problem was derived\nin [Bou17]. Later, [Mis21] detailed that expression to the case of planar motion of two\nellipsoids, which is given by\nV=\u0000GM 1M2\nrX\n(l1;m1)2\u0007\n(l2;m2)2\u0007\u0000l1;m1\nl2;m2\u0012R1\nr\u00132l1\u0012R2\nr\u00132l2\n\u0002\n\u0002Z1)\n2l1;2m1Z2)\n2l2;2m2cos(2m1(\u00122\u0000f) + 2m2(\u00122\u0000f));\nwhere\n\u0007 =f(l;m)2Z2: 0\u0014jmj\u0014lg;\nand, if we take L=l1+l2andM=m1+m2, the constants\n\u0000l1;m1\nl2;m2=(\u00001)L\u0000M\n4Lp\n(2l1\u00002m1)!(2l1+ 2m1)!(2l2\u00002m2)!(2l2+ 2m2)!(2L\u00002M)!(2L+ 2M)!\n(L\u0000M)!(L+M)!\nare numbers dealing with the interaction between the extended bodies. On the other\nhand,RjandZj)\n2lj;2mjare, respectively, the mean radius and the Stokes coe\u000ecients of\neachEj. The quantitiesZj)\nl;mprovide the expansion of the potential created for the body\nEj. They are related to the usual parameters Cj)\nl;mandSj)\nl;mby\nCj)\nl;m+iSj)\nl;m= (\u00001)m 2\n1 +\u000em;0s\n(l\u0000m)!\n(l+m)!\u0016Zj)\nl;m; m\u00150;\nwhere\u000em;nis the Kronecker delta.\nAppendix B.A different formulation of the equations of motion\nTo perform the integration of (12), it is convenient to adopt the eccentric anomaly u\nas independent variable, according to the following procedure.\nLet us write the equations (12) as\nCjd2\u0012j\ndt2(t) +\u0010a\nr\u00115\n\"jFj(t;\u0012) = 0; j = 1;2; (55)36 A. CELLETTI, J. GIMENO, AND M. MISQUERO\nwhere\u0012= (\u00121;\u00122),C1+C2= 1 andFj=@\u0012j(V2+V4),j= 1;2. The explicit expression\nforF1andF2is given by\nF1(t;\u0012) =\u0012\u0010r\na\u00112\n+5\n4\u0000\n^q2+5\n7^q1\u0001\u0013\nsin(2\u00121\u00002f)\n+25^d1\n8sin(4\u00121\u00004f) +3^d2\n8sin(2\u00121\u00002\u00122) +35^d2\n8sin(2\u00122+ 2\u00121\u00004f)\nF2(t;\u0012) =\u0012\u0010r\na\u00112\n+5\n4\u0000\n^q1+5\n7^q2\u0001\u0013\nsin(2\u00122\u00002f) +25^d2\n8sin(4\u00122\u00004f)\n\u00003^d1\n8sin(2\u00121\u00002\u00122) +35^d1\n8sin(2\u00122+ 2\u00121\u00004f): (56)\nLet us consider the change of variables given by the Kepler's equation\nxj(u) =\u0012j(u\u0000esinu); j = 1;2:\nThen, we have\nd2\u0012j\ndt2(t) =\u0012a\nr(u)\u00132d2xj\ndu2(u)\u0000\u0012a\nr(u)\u00133dxj\ndu(u)esinu; j = 1;2;\nso that, assuming Cj6= 0, (55) becomes\nd2xj\ndu2(u)\u0000a\nr(u)dxj\ndu(u)esinu+\u0012a\nr(u)\u00133\"j\nCjFj(u;x) = 0; j = 1;2: (57)\nWe can write the system (57) as\ndxj\ndu(u) =yj(u)\ndyj\ndu(u) =a\nr(u)yj(u)esinu\u0000\u0012a\nr(u)\u00133\"j\nCjFj(u;x);\nforj= 1;2 andx= (x1;x2). From the well-known relations used in the study of Kepler's\nproblem\ncosf=cosu\u0000e\n1\u0000ecosu; sinf=p\n1\u0000e2sinu\n1\u0000ecosu;\nwe can de\fne the functions s=s(xj),c=c(xj) as\ns(xj) = sin(2xj)(2 cos2f\u00001)\u0000cos(2xj)2 cosfsinf\nc(xj) = cos(2xj)(2 cos2f\u00001) + sin(2xj)2 cosfsinf ;THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 37\nso thatF1andF2in (56) take the form\nF1(u;x) =\u0012\u0012r(u)\na\u00132\n+5\n4\u0000\n^q2+5\n7^q1\u0001\u0013\ns(x1)\n+25^d1\n4s(x1)c(x1) +3^d2\n8sin(2x1\u00002x2) +35^d2\n4s(x1+x2)c(x1+x2)\nF2(u;x) =\u0012\u0012r(u)\na\u00132\n+5\n4\u0000\n^q1+5\n7^q2\u0001\u0013\ns(x2) +25^d2\n4s(x1)c(x1)\n\u00003^d1\n8sin(2x1\u00002x2) +35^d1\n4s(x1+x2)c(x1+x2):\nAppendix C.Expansion of V2andV4\nWe give below the expansion of V2andV4up to second order in the eccentricity:\nV2=\u00003d1e2GM 2cos(2\u00121)\n8a3\u00003d2e2GM 1cos(2\u00122)\n8a3\u000027d1e2GM 2cos(2t\u00002\u00121)\n8a3\n\u00003d1e2GM 2cos(4t\u00002\u00121)\n4a3\u000027d2e2GM 1cos(2t\u00002\u00122)\n8a3\u00003d2e2GM 1cos(4t\u00002\u00122)\n4a3\n\u00009d1eGM 2cos(t\u00002\u00121)\n8a3\u00009d1eGM 2cos(3t\u00002\u00121)\n8a3\u00009d2eGM 1cos(t\u00002\u00122)\n8a3\n\u00009d2eGM 1cos(3t\u00002\u00122)\n8a3\u00003d1GM 2cos(2t\u00002\u00121)\n4a3\u00003d2GM 1cos(2t\u00002\u00122)\n4a3\n\u00003e2GM 2q1cos(2t)\n8a3\u00003e2GM 1q2cos(2t)\n8a3\u00009e2GM 2q1\n8a3\u00009e2GM 1q2\n8a3\u00003eGM 2q1cos(t)\n4a3\n\u00003eGM 1q2cos(t)\n4a3\u0000GM 2q1\n4a3\u0000GM 1q2\n4a3;\nV4=\u0000225GM 2q2\n1e2\n112a5M1\u0000225Gcos(2t)M2q2\n1e2\n224a5M1\u0000225GM 1q2\n2e2\n112a5M2\u0000225Gcos(2t)M1q2\n2e2\n224a5M2\n\u0000105Gcos(2t\u00002\u00121\u00002\u00122)d1d2e2\n16a5\u0000525Gcos(4t\u00002\u00121\u00002\u00122)d1d2e2\n16a5\n\u0000315Gcos(6t\u00002\u00121\u00002\u00122)d1d2e2\n32a5\u000045Gcos(2\u00121\u00002\u00122)d1d2e2\n16a5\n\u000045Gcos(2t+ 2\u00121\u00002\u00122)d1d2e2\n64a5\u000045Gcos(2t\u00002\u00121+ 2\u00122)d1d2e2\n64a5\u0000225Gd2\n1M2e2\n224a5M1\n\u0000225Gcos(2t)d2\n1M2e2\n448a5M1\u000075Gcos(2t\u00004\u00121)d2\n1M2e2\n32a5M1\u0000375Gcos(4t\u00004\u00121)d2\n1M2e2\n32a5M1\n\u0000225Gcos(6t\u00004\u00121)d2\n1M2e2\n64a5M1\u000075Gcos(2t\u00002\u00122)d2q1e2\n8a5\u0000165Gcos(4t\u00002\u00122)d2q1e2\n64a5\n\u0000135Gcos(2\u00122)d2q1e2\n64a5\u0000375Gcos(2t\u00002\u00121)d1M2q1e2\n56a5M1\u0000825Gcos(4t\u00002\u00121)d1M2q1e2\n448a5M1\n\u0000675Gcos(2\u00121)d1M2q1e2\n448a5M1\u000075Gcos(2t\u00002\u00121)d1q2e2\n8a5\u0000165Gcos(4t\u00002\u00121)d1q2e2\n64a538 A. CELLETTI, J. GIMENO, AND M. MISQUERO\n\u0000135Gcos(2\u00121)d1q2e2\n64a5\u000045Gq1q2e2\n8a5\u000045Gcos(2t)q1q2e2\n16a5\u0000375Gcos(2t\u00002\u00122)d2M1q2e2\n56a5M2\n\u0000825Gcos(4t\u00002\u00122)d2M1q2e2\n448a5M2\u0000675Gcos(2\u00122)d2M1q2e2\n448a5M2\u0000225Gd2\n2M1e2\n224a5M2\u0000225Gcos(2t)d2\n2M1e2\n448a5M2\n\u000075Gcos(2t\u00004\u00122)d2\n2M1e2\n32a5M2\u0000375Gcos(4t\u00004\u00122)d2\n2M1e2\n32a5M2\u0000225Gcos(6t\u00004\u00122)d2\n2M1e2\n64a5M2\n\u0000225Gcos(t)M2q2\n1e\n224a5M1\u0000225Gcos(t)M1q2\n2e\n224a5M2\u0000525Gcos(3t\u00002\u00121\u00002\u00122)d1d2e\n64a5\n\u0000525Gcos(5t\u00002\u00121\u00002\u00122)d1d2e\n64a5\u000045Gcos(t+ 2\u00121\u00002\u00122)d1d2e\n64a5\n\u000045Gcos(t\u00002\u00121+ 2\u00122)d1d2e\n64a5\u0000225Gcos(t)d2\n1M2e\n448a5M1\u0000375Gcos(3t\u00004\u00121)d2\n1M2e\n128a5M1\n\u0000375Gcos(5t\u00004\u00121)d2\n1M2e\n128a5M1\u000075Gcos(t\u00002\u00122)d2q1e\n32a5\u000075Gcos(3t\u00002\u00122)d2q1e\n32a5\n\u0000375Gcos(t\u00002\u00121)d1M2q1e\n224a5M1\u0000375Gcos(3t\u00002\u00121)d1M2q1e\n224a5M1\u000075Gcos(t\u00002\u00121)d1q2e\n32a5\n\u000075Gcos(3t\u00002\u00121)d1q2e\n32a5\u000045Gcos(t)q1q2e\n16a5\u0000375Gcos(t\u00002\u00122)d2M1q2e\n224a5M2\n\u0000375Gcos(3t\u00002\u00122)d2M1q2e\n224a5M2\u0000225Gcos(t)d2\n2M1e\n448a5M2\u0000375Gcos(3t\u00004\u00122)d2\n2M1e\n128a5M2\n\u0000375Gcos(5t\u00004\u00122)d2\n2M1e\n128a5M2\u000045GM 2q2\n1\n224a5M1\u000045GM 1q2\n2\n224a5M2\u0000105Gcos(4t\u00002\u00121\u00002\u00122)d1d2\n32a5\n\u00009Gcos(2\u00121\u00002\u00122)d1d2\n32a5\u000045Gd2\n1M2\n448a5M1\u000075Gcos(4t\u00004\u00121)d2\n1M2\n64a5M1\n\u000015Gcos(2t\u00002\u00122)d2q1\n16a5\u000075Gcos(2t\u00002\u00121)d1M2q1\n112a5M1\n\u000015Gcos(2t\u00002\u00121)d1q2\n16a5\u00009Gq1q2\n16a5\u000075Gcos(2t\u00002\u00122)d2M1q2\n112a5M2\u000045Gd2\n2M1\n448a5M2\n\u000075Gcos(4t\u00004\u00122)d2\n2M1\n64a5M2:\nReferences\n[Arn63] V. I. Arnol'd. Proof of a theorem of A. N. Kolmogorov on the invariance of quasi-periodic\nmotions under small perturbations. Russian Math. Surveys , 18(5):9{36, 1963.\n[Bel66] V. V. Beletskii. Motion of an arti\fcial satellite about its center of mass . Mechanics of\nSpace Flight. Israel Program for Scienti\fc Translations; [available from the U.S. Dept. of\nCommerce, Clearinghouse for Federal Scienti\fc and Technical Information, Spring\feld,\nVa.], Jerusalem, 1966.\n[BL75] V. V. Beletskii and E. K. Lavrovskii. On the theory of the resonance rotation of Mercury.\nAstronomicheskii Zhurnal , 52:1299{1308, December 1975.\n[BL09] Gwena el Bou\u0013 e and Jacques Laskar. Spin axis evolution of two interacting bodies. Icarus ,\n201(2):750 { 767, 2009.THE SPIN-SPIN PROBLEM IN CELESTIAL MECHANICS 39\n[BM15] Konstantin Batygin and Alessandro Morbidelli. Spin-Spin coupling in the Solar System.\nThe Astrophysical Journal , 810(2):110, sep 2015.\n[BO16] Alberto Boscaggin and Rafael Ortega. Periodic solutions of a perturbed kepler problem\nin the plane: From existence to stability. Journal of Di\u000berential Equations , 261(4):2528{\n2551, 2016.\n[Bou17] Gwena el Bou\u0013 e. The two rigid body interaction using angular momentum theory formulae.\nCelestial Mechanics and Dynamical Astronomy , 128(2):261{273, Jun 2017.\n[CC00] Alessandra Celletti and Luigi Chierchia. Hamiltonian stability of spin{orbit resonances in\ncelestial mechanics. Celestial Mechanics and Dynamical Astronomy , 76:229{240, 05 2000.\n[CC08] A. Celletti and L. Chierchia. Measures of basins of attraction in spin-orbit dynamics.\nCelestial Mechanics and Dynamical Astronomy , 101:159{170, May 2008.\n[CC09] A. Celletti and L. Chierchia. Quasi-Periodic Attractors in Celestial Mechanics. Archive\nfor Rational Mechanics and Analysis , 191:311{345, February 2009.\n[CCGdlL21a] Renato C. Calleja, Alessandra Celletti, Joan Gimeno, and Rafael de la Llave. Break-down\nthreshold of invariant attractors in the dissipative spin-orbit problem. Preprint , 2021.\n[CCGdlL21b] Renato C. Calleja, Alessandra Celletti, Joan Gimeno, and Rafael de la Llave. E\u000ecient\nand accurate KAM tori construction for the dissipative spin-orbit problem using a map\nreduction. Preprint , 2021.\n[CCGdlL21c] Renato C. Calleja, Alessandra Celletti, Joan Gimeno, and Rafael de la Llave. KAM quasi-\nperiodic tori for the dissipative spin-orbit problem. Preprint , 2021.\n[Cel90] A. Celletti. Analysis of resonances in the spin-orbit problem in celestial mechanics: The\nsynchronous resonance (Part I). Zeitschrift Angewandte Mathematik und Physik , 41:174{\n204, March 1990.\n[Cel10] Alessandra Celletti. Stability and Chaos in Celestial Mechanics . Springer, Berlin, Heidel-\nberg, 01 2010.\n[CFL04] A. Celletti, C. Falcolini, and U. Locatelli. On the break-down threshold of invariant tori\nin four dimensional maps. Regul. Chaotic Dyn. , 9(3):227{253, 2004.\n[DeV58] Ren\u0013 e DeVogelaere. On the Structure of Symmetric Periodic Solutions of Conservative\nSystems, with Applications , pages 53{84. Princeton University Press, 1958.\n[Eke90] I. Ekeland. Convexity Methods in Hamiltonian Mechanics , volume 19 of Ergebnisse der\nMathematik und ihrer Grenzgebiete : a series of modern surveys in mathematics. Folge\n3. Springer-Verlag, Berlin Heidelberg, 1990.\n[Gol80] Herbert Goldstein. Classical Mechanics . Addison-Wesley, 1980.\n[GP66] Peter Goldreich and Stanton Peale. Spin orbit coupling in the Solar System. The Astro-\nnomical Journal , 71:425, 07 1966.\n[Gre79] John M. Greene. A method for determining a stochastic transition. J. Math. Phys. ,\n20(6):1183{1201, 1979.\n[HX17] Xiyun Hou and Xiaosheng Xin. A note on the spin{orbit, spin{spin, and spin{orbit{spin\nresonances in the binary minor planet system. The Astronomical Journal , 154(6):257, nov\n2017.\n[HZ99] Alan Hampton and Dami\u0013 an H. Zanette. Measure synchronization in coupled hamiltonian\nsystems. Phys. Rev. Lett. , 83:2179{2182, Sep 1999.\n[JA16] Mahdi Jafari Nadoushan and Nima Assadian. Geography of the rotational resonances and\ntheir stability in the ellipsoidal full two body problem. Icarus , 265:175 { 186, 2016.\n[JNA16] Mahdi Jafari Nadoushan and Nima Assadian. Chirikov di\u000busion in the sphere-ellipsoid\nbinary asteroids. Nonlinear Dynamics , 85, 08 2016.\n[JZ05] \u0012A. Jorba and M. Zou. A software package for the numerical integration of ODEs by means\nof high-order Taylor methods. Experiment. Math. , 14(1):99{117, 2005.\n[Kin72] H. Kinoshita. First-order perturbations of the two \fnite body problem. Publ. Astron. Soc.\nJapan , 24:423, January 1972.\n[Kol54] Andrej N. Kolmogorov. On the conservation of conditionally periodic motions under small\nperturbation of the hamiltonian. In Dokl. Akad. Nauk. SSR , volume 98, pages 2{3, 1954.\n[Kre50] M.G. Krein. Generalization of certain investigations of A.M. Lyapunov on linear di\u000ber-\nential equations with periodic coe\u000ecients. Dokl. Akad. Nauk USSR , 73:445{448, 1950.40 A. CELLETTI, J. GIMENO, AND M. MISQUERO\n[LR98] Jeroen S.W. Lamb and John A.G. Roberts. Time-reversal symmetry in dynamical sys-\ntems: A survey. Physica D: Nonlinear Phenomena , 112(1):1{39, 1998. Proceedings of the\nWorkshop on Time-Reversal Symmetry in Dynamical Systems.\n[Mac95] Andrzej J. Maciejewski. Reduction, relative equilibria and potential in the two rigid bodies\nproblem. Celestial Mechanics and Dynamical Astronomy , 63(1):1{28, Mar 1995.\n[Mis21] Mauricio Misquero. The spin-spin model and the capture into the double synchronous\nresonance. Nonlinearity , 34:2191{2219, 2021.\n[MO20] Mauricio Misquero and Rafael Ortega. Some rigorous results on the 1:1 resonance of the\nspin-orbit problem. SIAM J. Applied Dynamical Systems , 19(4):2233{2267, 2020.\n[Mos62] J. Moser. On invariant curves of area-preserving mappings of an annulus. Nachr. Akad.\nWiss. G ottingen Math.-Phys. Kl. II , 1962:1{20, 1962.\n[MW79] W. Magnus and S. Winkler. Hill's Equation . Dover, New York, 1979.\n[Sch72] H.R. Schwarz. Stability of Kepler motion. Computer Methods in Applied Mechanics and\nEngineering , 1(3):279{299, 1972.\n[Sch02] D. J. Scheeres. Stability in the Full Two-Body Problem. Celestial Mechanics and Dynam-\nical Astronomy , 83(1):155{169, 2002.\n[Sch09] D. J. Scheeres. Stability of the planar full 2-body problem. Celestial Mechanics and Dy-\nnamical Astronomy , 104(1):103{128, Jun 2009.\n[Ver78] J.H. Verner. Explicit Runge-Kutta methods with estimates of the local truncation error.\nSIAM J. Numer. Anal. , 15(4):772{790, 1978.\n[YS75] V. A. Yakubovich and V. M. Starzhinskii. Linear Di\u000berential Equations with Periodic\nCoe\u000ecients . Wiley, New York, 1975.\n[ZOST66] V.A. Zlatoustov, D.E. Ohotzimsky, V.A. Sarychev, and A.P. Torzhevsky. Investigation of\na satellite oscillations in the plane of an elliptic orbit. In G ortler H., editor, Applied Me-\nchanics. Proceedings of the Eleventh International Congress of Applied Mechanics Munich\n(Germany) 1964 , pages 436{439. Springer, Berlin, Heidelberg, 1966.\nDepartment of Mathematics, University of Rome Tor Vergata, Via della Ricerca\nScientifica 1, 00133 Rome (Italy)\nEmail address :celletti@mat.uniroma2.it\nDepartment of Mathematics, University of Rome Tor Vergata, Via della Ricerca\nScientifica 1, 00133 Rome (Italy)\nEmail address :gimeno@mat.uniroma2.it\nDepartment of Mathematics, University of Rome Tor Vergata, Via della Ricerca\nScientifica 1, 00133 Rome (Italy)\nEmail address :misquero@mat.uniroma2.it" }, { "title": "1211.4926v1.The_quantum_dynamics_of_two_coupled_large_spins.pdf", "content": "The quantum dynamics of two coupled large spins \nV.E. Zobov \nL.V.Kirensky Institu te of Physics, Ru ssian Academy of Sciences, Siberian Branch, \n660036,Krasnoyarsk, R ussia; rsa@iph.krasn.ru \n \nWe calculate the tim e evolution of m ean spin com ponents and the squared I-concurrence \nof two coupled large spins S. As the initial conditions we take tw o cases: spin coherent states and \nuniform superposition states. For the spin cohere nt states we have obtained the asym ptotic for-\nmulas at S>>1 and t<> 1, t <>1 for the squared I-concu rrence can be obtained a \nsimple asymptotic form ula. In Eq. (9 ) for when t<>1, thes e minima can be de-\nscribed by G aussian functions as: \n∑=\n= ⎪⎭⎪⎬⎫\n⎪⎩⎪⎨⎧\n⎟⎠⎞⎜⎝⎛− −≈⎥⎦⎤\n⎢⎣⎡\n⎭⎬⎫\n⎩⎨⎧Mn\nnS\nMnTtJSM\nSJtM\n02\n22 4\n2 2exp2cos . \nThen to get a squared I-concurrence \n⎟⎟\n⎠⎞\n⎪⎭⎪⎬⎫\n⎪⎩⎪⎨⎧\n⎥⎥\n⎦⎤\n⎢⎢\n⎣⎡\n⎟\n⎠⎞⎜\n⎝⎛−+− −⎜⎜\n⎝⎛\n−−+−\n+−+=\n∑∑=\n==\n=Mn\nnS M\nMcoh\nMnTtJSM\nSS StJerf\ntJ SSC\n12\n22 2\n122\n222\n212exp\n2221)811(\n11121 2\nππ\n. (13) \nIn Fig. 3 perfor med a comparison of the de pendences, calculated on the exact and asymp-\ntotic f ormulas. In the las t (13), due to the rap id decrease of amplitude with increasing M, we are \nrestricted by the first three i ngredients with M = 2, 3 and 4. We see good agreem ent the asym p-\ntotic form ula with exac t one. W hen S increas es its accu racy will in crease, and the dep th of m in-\nima will be r educed. \nThe changes in behaviors and with the growth of S shown in Fig. 4. The \ntime evolution of slows down and thus closes in th e motion of classical magnetic m o-\nments. W hile the squared I-conc urrence rea ches m aximum value C = 1, indicating preservation \nof the quantum properties. Previously, such a dependence of entanglem ent has been dem on-)(t Fcoh )(2t Ccoh\n)(t Fcoh\n 5strated in another m odel system : a particle with spin, oscillating in a non-uniform magnetic field \n[8]. \n \nConclusions \nIf the two sp ins are p repared in the sp in coherent states, the m otion of their m ean projec-\ntions in the lim it S → ∞ will conc ede with the m otion of the p rojections of two class ical mag-\nnetic m oments. Neverth eless, the sq uared I -conc urrenc e, and thus the entanglem ent will grow \nwith tim e to values app roaching to th e maximum C = 1 when S → ∞. \nIf the two sp ins are p repared in the un iform super position states, we observe the quantum \ndynam ics in any S, so far as classical m agnetic mo ments can not be in supe rposition state (like of \nthe fam ous «Schrodinger cat» [14] ). (Although the preparation of th e superposition state at large \nS can be technically very difficult). \nThus, the large spins m ay show both the quant um and classical prop erties, accord ing to \nprepared states and observe conditions. It is im portant for us (and this is the m ain goal of this \nstudy) that quantum computation on qudits, represented by large sp ins, do not get the principal \nlimitations o n the spin n umber S. \n \nThis work was supported by the R ussian Founda tion for Basic Research (project. no. 09-07-\n00138-a). \n \n 6References \n1. D. Gottesm an,.Lect. Not. Co mp. Sci. 1509, 302 (1999). \n2. A.R. Kessel, V.L. Er makov, JETP Letters 70, 61 (1999); 71, 307 (2000). \n3. A. Muthukrishnan, C. R. Stroud (Jr.), Phys. Rev. A 62, 052309 (2000). \n4. B.C. Sanders, Phys. Rev. A 40, 2417 (1989). \n5. J.L. Garcia-Palacios, S Dattagupta, Phys. Rev. Lett. 95, 190401 (2005). \n6. A. Polkovnikov, “Representation of quantum dynamics of interacting system s through clas-\nsical trajectories ”, cond-m at/0905.3384. \n7. A Cabello, Phys. Rev. A 65, 062105 (2002). \n8. S. Ghose, P.M. Alsing, B.C. Sanders, I.H. Deutsch, Phys. Rev. A 72, 014102 (2005). \n9. G. Burlak, I. Sainz, A.B. Klim ov, Phy s. Rev. A 80, 024301 (2009). \n10. P. Rungta, V. Buzek, C.V. Caves et al., Phys. Rev. A 64, 042315 (2001); P. Rungta, C.V. \nCaves, Phys. Rev. A 67, 012307 (2003). \n11. Y.Yang, W .Liu, Z. Sun, X. W ang, Phys. Rev. A 79, 054104 (2009). \n12. X. Lu, X. Wang, Y.Yang, J. Chen, Quantum Inf. Co mput. 8, 671 (2008); quant-\nph/0803.1032. \n13. A.S. Er milov, V.E. Zob ov, Opt. Spectrosc. 103, 969 (2007). \n14. R. Penrose, «The Em peror' s New Mind», Oxford University Press, 1989. \n 7Квантовая динамика двух взаимодействующих больших спинов \nВладимир Евгеньевич Зобов \nИнститут физики им. Л. В. Киренского СО РАН , \n660036, Красноярск , Академгород ок,50, стр.38. \nrsa@iph.krasn.ru\n \n Рассчитывается временная эволюция средних спиновых компонент и квадратичной \nI-согласованности двух взаимодейству ющих больших спинов S. В качестве начальных ус-\nловий взяты два случая : спиновые когерентные состояния и равномерные суперпозицион -\nные состояния . Для спиновых когерентных состояний получено асимптотическое выра -\nжение при S>>1 и t< 0 (Fig. 1e) at high |E|. Tuning |E| to approximately 7 kV/cm where c = 0 would realize the Heisenberg XXX model. In future studies, such tunability will allow exploration of qualitatively different spin dynamics hosted by a spin-1/2 system16. In 2D, the average dipolar interaction can be controlled by orienting the dipole moments relative 6 to the 2D plane, as a direct consequence of the anisotropy of the dipolar interaction. We measure c between |¯ñ and |1ñ as a function of a at |E| = 1.02 kV/cm. By rotating the dipoles, c is varied continuously from negative to positive (Fig. 2g). At a = 0°, molecules interact repulsively as the dipole moments are aligned perpendicular to the 2D plane, leading to c < 0. At a = 90°, dipoles are aligned within the plane and the average interaction is attractive, leading to c > 0. Assuming symmetric transverse trapping, c µ -(3cos2a - 1). Notably, this tunability of c by rotating E arises from a common geometric factor on the Ising and exchange interactions, in contrast to tuning with |E|, which controls the relative strength of the two interactions. The choice of the internal states used for the spin-1/2 system offers additional control over the interaction parameters. The green squares in Fig. 2e show a measurement of Uc between |¯ñ and |2ñ = |1, -1ñ at |E| = 0. Using |2ñ instead of |1ñ changes the magnitude of c and reverses its sign at |E| = 0, as a result of the different transition dipole moments in {|¯ñ, |2ñ} where 𝑑↑↓=−𝑑↓↑=⟨↓|𝑑|↑/⟩/√2. We measure c between |¯ñ and |2ñ to be 𝜒0 = 2.3(8)´10-6 Hz/cm-2, consistent with -1/2 of the interaction between |¯ñ and |1ñ at this field. The ability to manipulate dipolar interactions through the internal state allows us to dynamically change the interaction and the evolution of the many-body state. In contrast to tuning with the dc electric field, internal state control of the interaction can be achieved rapidly and coherently with microwave pulses. Such a capability is essential for applications such as dynamic Hamiltonian engineering35, high fidelity quantum gate operations36, and studying quantum quenched dynamics with dipolar spin systems37,38. By implementing a dynamical control pulse sequence, we demonstrate reversal of the coherent many-body spin dynamics. In particular, we coherently swap the excited state from |2ñ to |1ñ in the middle of the Ramsey evolution, which instantaneously changes the interaction parameter c by a factor of -2, and consequently reverses the associated phase progress in a coherent evolution. Our sequence consists of two phase accumulation stages (Fig. 3a). After preparation in a superposition of |¯ñ, |2ñ, molecules evolve with 𝜒0 (stage I) until a composite microwave pulse R (Inset of Fig. 3a) coherently transfers the excited state population from |2ñ to |1ñ while 7 maintaining the phase relative to |¯ñ. Molecules then continue to evolve in {|¯ñ, |1ñ} (stage II) with 𝜒/=−2𝜒0 before being detected after the final Ramsey pulse. The XY8 decoupling sequence is used during both stages. The duration of R is 70 µs, which is short compared to the interaction timescales. We then extract the phase shift DF = Df(q = p/4) -Df(q = 3p/4) as a function of T, plotted in Fig. 3b. (see Methods) We observe a sign change in the rate of phase accumulation d(DF)/dT at ts = 1.2 ms when R is implemented, indicating a reversal of the mean-field interaction sign. The data is well described by a piecewise linear fit with the ratio of d(DF)/dT before and after the reversal constrained to be −2, consistent with the measurement in Fig. 2e. At T » 1.6 ms, DF returns to the initial value, completing the phase reversal. Our protocol realizes a complete reversal of the spin Hamiltonian at |E| = 0. Such a capability is essential for studying out-of-time-ordered correlators which are used to understand dynamics of interacting quantum many-body systems such as quantum information scrambling39-42. Similar many-body echo processes also allow applications such as robust Heisenberg-limited phase sensitivity without single-particle-resolved state detection27,43-45. At long evolution times, motional and spin dynamics are coupled by dipolar interactions, resulting in dynamical evolution beyond the spin model. Mode-changing dipolar collisions lead to loss of coherence, manifesting as an exponential decay of the Ramsey contrast, exp(-G t). G is proportional to the rate of elastic collisions ns vT, where vT is the thermal velocity and s is the cross section for dipolar elastic collisions. For spin-polarized fermionic molecules with dipole moment d induced by E, previous theories46,47 and experiments22,24 have shown s µ d4. By comparison, two molecules in a coherent superposition carry oscillating dipoles with amplitudes of 𝑑↓↑ in the lab frame, leading to dipolar collisions even at low |E| where induced dipole moments are small26,48. When the oscillations are fully coherent, the molecules collide with maximum effective dipole moment 𝑑↓↑/√2 and s (Ref. 48). These collisions couple the spin and motion, resulting in spin decoherence that cannot be removed using multi-pulse sequences, in contrast to interaction-induced dephasing in a lattice13. We study the effect of dipolar collisions by measuring contrast decay as a function of n. To 8 investigate collisional decoherence, the single particle dephasing rate must be below the collisional rate, which was achieved by increasing the number of decoupling pulses. We observe the characteristic exponential decay of the contrast for a range of n (Fig. 4a). The extracted G increases with density (Fig. 4b), indicating collision-limited coherence time. At n = 1.1(1) ´ 107 cm-2 and |E| = 0, we measure G = 130(5) s-1, consistent with dipolar elastic collision rates measured in Ref. 22 with d being 𝑑↓↑/√2 (see Methods). This gives a collisional timescale shorter than the trap oscillation period of ~22 ms (see Methods). Rotating the electric field to a = 36° at |E| = 1.02 kV/cm reduces the strength of the dipolar interaction. We observe longer coherence times at the same density (purple squares in Fig. 4b), indicating reduced s. By extracting the average s = G/nvT for each E, we found s(a = 0°, |E| = 0)/s(36°, 1.02 kV/cm) = 3.1(4). This variation of s is expected to arise from the changing dipolar cross section as a function of a, similar to scattering with induced dipole moments24,47. In analogy to the relationship in atomic systems between the mean-field shift and elastic cross section, the variation of s could also be explained by a scaling s µ c2. The measured ratio of s is consistent with [c(0°, 0)/c(36°, 1.02 kV/cm)]2 = 3.5(9) from Fig. 2f, 2g. These results suggest that dipolar effects dominate the collisional decoherence process. The high controllability of our system allows tuning the motional dynamics relative to the coherent spin dynamics by changing parameters such as the strength of optical trapping49, temperature, and dipolar interaction strength. These capabilities would allow the investigation of other many-body spin-motion effects such as unconventional paired superfluidity50,51, spin-wave and spin transport physics52,53, and dipole-mediated spin-orbital dynamics54. In conclusion, we have demonstrated tunable dynamics of the collective spin and motion in a 2D itinerant spin system formed from a gas of fermionic polar molecules. We have shown static and dynamical control over the spin Hamiltonian using external fields, microwaves, and internal states of the molecules. Our results establish a highly controllable spin system that can be used to investigate a broad range of many-body phenomena. 9 Methods Molecular spin system and preparation We create ultracold gases of KRb in 2D with the procedure detailed in Ref. 22 and Ref. 28. In brief, degenerate gases of 40K and 87Rb are loaded into a stack of 2D harmonic traps formed by a one-dimensional optical lattice. Molecules in the ground rotational state |¯ñ are subsequently produced by magneto-association and stimulated Raman adiabatic passage (STIRAP) directly at the electric field E. We typically produce 15´103 molecules at a temperature T0 » 450 nK, occupying 19(1) layers with a peak number of about 1000 molecules in a single layer. The harmonic trapping frequencies within each layer for molecules in |¯ñ at |E| = 0 are (wx, wy, wz) = 2p ´ (45, 17000, 45) Hz. Since the harmonic level spacing ħwy along the tight confinement direction is larger than kBT0 and the interaction energy scale, our molecules predominantly remain in the lowest harmonic level, forming a quasi-2D system throughout the experiments. Here, kB is the Boltzmann constant. Couplings between these 2D layers are weak compared to the intralayer interaction28 due to the large spatial extension of the fermionic molecular gases in each layer that averages out the interlayer interaction. This weak interlayer interaction is expected to manifest as only a small correction to c for the timescales considered in this work. We therefore approximate the system as a stack of isolated layers. At |E| = 0 and typical trapping condition, the resonant transition frequency between |¯ñ and |1ñ (|¯ñ and |2ñ ) is measured to be around 2228.138 MHz (2227.742 MHz). Compared to our typical Rabi frequency of 100 kHz, the large difference between resonant frequencies allows us to resolve these states by changing the microwave frequency. The degeneracy of |1,1ñ and |2ñ = |1, -1ñ is broken by a weak coupling of the nuclear and rotational degrees of freedom, resulting in an energy difference of about 100 kHz between the two states. To avoid off-resonant transfer to |1,1ñ when driving the transition between |¯ñ and |2ñ, we reduce the microwave Rabi frequency to about 60 kHz. Besides rotational structure, molecules possess hyperfine states. In our experiment, KRb is prepared in |¯ñ = |N = 0, mN = 0, mK = -4, mRb = 1/2ñ. Here, mK and mRb are the projection of the 10 nuclear spins on the quantization axis. For states with N = 1 (|1ñ and |2ñ), electric quadrupole interaction slightly mixes the hyperfine states and the rotational states. Since hyperfine changing transitions are weak and their energy spectrum is well resolved, we use states with the largest projection on |1, 0, -4, 1/2> and |1, -1, -4, 1/2> as |1ñ and |2ñ, respectively. In this work, we avoid driving hyperfine-changing transitions, and thus neglect the hyperfine degree of freedom. The density of the molecular gas is varied by adding an extra cleaning procedure after the molecules are produced and before the Ramsey interrogation. Specifically, a microwave pulse with an area of qc prepares molecules in a superposition of |¯ñ and |1ñ. The STIRAP beam that is resonant with |¯ñ and the electronically-excited intermediate state of the molecules is switched on for 50 µs to eliminate the |¯ñ component of the superposition. The remaining molecules are then flipped back to |¯ñ with a p pulse. The density is reduced to sin2(qc /2) of the initial density. At certain values of E, the cleaning STIRAP light pulse can also off-resonantly excite molecules in |1ñ. In that case, molecules in |1ñ are shelved in |2,0ñ during the optical blast using an additional microwave pulse, and no loss of |1ñ molecules is observed. Since the size of the STIRAP beam used for cleaning is larger than the size of the sample, we do not observe significant temperature change from the process. This method therefore allows us to control the sample density without changing the temperature or the spatial distribution in the regime where the experiments are conducted. Timescales of the dynamics Multiple timescales are involved in the dynamics presented in this work: 1) interaction timescale set by Uc, 2) collision timescale set by G , and 3) single-particle motion timescale set by the period of harmonic trapping. The spin model is valid in the regime t < 1/G (short time limit). For n = 1.1(1) ´ 107 cm-2 (where the strongest collisional effects are explored in the current work), 1/G » 7.4 ms. At this density, Uc » 60 Hz, giving a timescale of 1/ Uc » 17 ms. Since Uc µ 𝜒 while s, therefore G, scales with c2, it is possible to tune the relative timescale of these two processes. Spin dynamics are favored for small c, while collisional dynamics dominate at large c. 11 The period of trap oscillation is ~22 ms. Therefore, for the highest density (1/G » 7.4 ms), molecules collide before they complete one trap oscillation (akin to the hydrodynamic regime). For the lowest collision rate we measured when we lowered the density or reduced the strength of dipolar interaction, molecules would undergo several oscillations before a collision occurs (collisionless regime). The relative rates of the two processes can be controlled by changing the strength of the optical trapping, temperature, density, or c. The capability of tuning the system across different regimes highlights the controllability of our system. 2D density calibration We use the following procedure to obtain the average 2D densities in the layers from the total molecule number Ntot. The molecules occupy multiple layers formed by an optical lattice. We first measure the molecule distribution using layer-resolved spectroscopy28. In brief, an electric field gradient along the longitudinal direction of the optical lattice shifts the transition frequency between |0,0ñ and |1,0ñ on each layer. When the differential shift for molecules in adjacent layers is large compared to the linewidth of the transition in each layer, molecules in different layers can be addressed, resolved, and measured spectroscopically using microwaves. A typical measurement (Extended data figure 1) reveals a Gaussian distribution with a full-width-half-maximum of 11(1) layers. \n Extended data figure 1. A typical molecular number distribution measured via layer-resolved spectroscopy. \n 12 For two-body processes considered in this work, such as the mean-field frequency shift and collisional decoherence, we calculate the average number of molecules per layer using an effective number of layers L, defined as\t𝐿=\tef∑egfg, where Nk is the number for molecules in the k-th layer. From the layer-resolved number distribution, we extract L = 19(1) and use N2D = Ntot/L as the average number of molecules in a single layer. The average 2D density n is calculated from N2D using temperatures measured with time-of-flight thermometry and the transverse trapping frequencies for |¯ñ at E. Our procedure for varying the density is not expected to change the number distribution across layers. We therefore use L = 19 for all our measurements. Geometric factor of c The dipolar interaction between two dipoles has a geometric factor (1-3cos2Qij), where Qij is the angle between the dipole moment and the vector between the two particles. When the molecular motion is confined to 2D, cos2Qij can be decomposed into a and the azimuth angle b as cos2Qij = sin2a cos2b. This geometric factor can be expressed in terms of a by taking the average over the harmonic oscillator states in the 2D plane. In the mean-field regime and with approximately symmetric transverse trapping, this averaging reduces to an average over b, yielding ácos2bñ = 1/2 and thus á1-3cos2Qijñ = (3cos2a -1)/2. Measurement of Ramsey coherence time For the coherence time measurements, we measure the contrast at different Ramsey time T with the number of XY8 echo pulses fixed. To extract the contrast at each T, we perform 8 to 16 measurements of the fraction of |1ñ with f equally spaced between 0 and 360 degrees. The contrast and its standard deviation (SD) at time T are extracted from the measured fractions via bootstrapping following the procedure described in Ref. 28. Dynamical decoupling and calibration of the pulse sequence Grey circles are the experimental measurements. Black solid line is a fit to a summation of equally spaced Gaussian functions with a global Gaussian envelope. The width of each narrow Gaussian peak is assumed to be the same. The data is taken at |E| = 1.02 kV/cm and a = 36° (magic condition) to reduce broadening of the single-layer transition linewidth due to differential polarizability. 13 We perform dynamical decoupling with an XY8 multi-pulse sequence to suppress single particle dephasing. The consecutive spin echo pulses spaced by time t form a bandpass filter that rejects noise outside of a window with center frequency f0 = 1/(2t) and width ~1/T, suppressing the Ramsey phase fluctuations caused by noise in E and the nonzero differential ac polarizability. The XY8 sequence’s time-reversal symmetry and alternation between rotation axes further improve robustness against pulse area error and finite pulse duration. Multiple XY8 sequences are concatenated to achieve longer Ramsey time T without affecting f0. The efficacy of the dephasing suppression depends on t, pulse duration tp, and the timing of the sequence, which we optimize with the following procedure: An ideal XY8 sequence consists of infinitely fast echo pulses. However, if the microwave Rabi frequency is too high, molecular hyperfine states that are not in the spin-1/2 manifold can be coupled off-resonantly by the microwave field. To avoid this effect, we limit the Rabi frequency such that the p pulses time is around 10 µs for measurements in {|¯ñ, |1ñ}, and around 16 µs for measurements in {|¯ñ, |2ñ}. At fixed microwave power, we precisely determine the pulse duration in order to reduce the rotation error of the pulses. To do this, we perform a p/2 rotation followed immediately by up to eight closely spaced echo pulses. We fine tune the pulse duration tp until equal population of the two spin states are obtained regardless of the number of echo pulses applied. The pulse spacing t determines the passband frequency of the noise filter. Shorter t gives higher passband frequencies and less sensitivity to low frequency electric field noise, which is the dominant noise source in our system. We characterize noise rejection as a function of t by measuring the Ramsey phase fluctuations after a fixed T with {|¯ñ, |1ñ}. We vary t by changing the number of equally spaced Rabi-p pulses inserted between the two Ramsey pulses. We choose the phase of the second Ramsey pulse to maximize the sensitivity of the fraction of |1ñ to the Ramsey phase. For each t, we repeat the measurement 10 times to extract the SD of Ramsey phase, as computed from the measured fraction of |1ñ. We use echoes along 𝑋b only for this experiment, which provides the same electric field noise rejection performance as XY8. 14 Extended data figure 2 shows the SD of Ramsey phase as a function of t for T = 1.5 ms. We observe reduction of the phase fluctuations as t is reduced. To balance between decoupling efficacy and a hight/tp ratio to minimize system evolution during the pulses, we chose t = 140 µs for the mean-field frequency shift measurements. To account for the finite widths of the Ramsey pulses, we adjusted the delay (t/2 for infinitely short pulses) after the first Ramsey pulse and before the final detection pulse to minimize the sensitivity of the Ramsey phase shift to detuning D at resonance (𝜕Df/𝜕D)D\tij and maximize the range of D within which the phase shift is insensitive to D at first order. The amount of the adjustment is determined by simulating the XY8 sequence on a two-level system, and depends on the rotation angle q, which is accounted for in the measurements. Extracting phase shifts for the reversal measurements Switching between two spin manifolds requires changing the frequency of the microwave source. We use two procedures to reduce extra phase shift introduced by this process: (1) we measure the Extended data figure 2. Noise suppression of the dynamical decoupling sequence. Data is taken at |E| = 1.02 kV/cm and a = 0°. We use X-echo only, which provides similar performance to XY8 in terms of rejecting noise in D. Each data point is extracted from 10 repetitions. The dashed line is the noise floor of our measurement. \n 15 phase difference DF =Df(q = p/4) - Df(q = 3p/4) between the Ramsey fringes obtained for q =p /4 and q =3p /4; (2) For q =3p /4 measurement, we first prepare the molecules in |2ñ instead of |¯ñ and then apply a q =p /4 pulse. This allows us to keep the timings of the sequences for q =p /4 and q =3p /4 identical. Dipolar collisions with transition dipole moment In Ref. 22, the dipolar elastic collision rate between spin polarized KRb in 2D was measured. With n » 5.0´107 cm-2, T0 » 250 nK, d = 0.2 D, Ref. 22 reported G 0 = 168(48) s-1. Scaling G 0 as 𝛤∝𝑛\tl𝑇j\t𝑑m with n = 1.1(1)´107 cm-2, T0 = 463(9) nK, and 𝑑=\t𝑑↓↑/√2\t(conditions for the grey circles in Fig. 4a), we obtain G = 95(29) s-1. Acknowledgement We thank Thomas Bilitewski and Ana Maria Rey for inspirational discussions and critical reading of the manuscript. We acknowledge funding from the DOE Quantum System Accelerator, the National Science Foundation QLCI OMA-2016244, the National Science Foundation Phys-1734006, and the National Institute of Standards and Technology. J. S. H. acknowledges support from the National Research Council postdoctoral fellowship. C.M. acknowledges the NDSEG Graduate Fellowship. Author contributions All authors performed the experiments, analyzed the data, and contributed to interpreting the results and writing the manuscript. Data availability The datasets generated and analyzed during the current study are available from the corresponding authors on reasonable request. Competing interests The authors declare no competing interests. 16 References 1. Manousakis, E. The spin-1/2 Heisenberg antiferromagnet on a square lattice and its application to the cuprous oxides. Rev. Mod. Phys. 63, 1–62 (1991). 2. Žutić, I., Fabian, J. & Das Sarma, S. Spintronics: Fundamentals and applications. Rev. Mod. Phys. 76, 323–410 (2004). 3. Wineland, D. J., Bollinger, J. J., Itano, W. M., Moore, F. L. & Heinzen, D. J. Spin squeezing and reduced quantum noise in spectroscopy. Phys. Rev. A 46, R6797--R6800 (1992). 4. Nielsen, M. A. & Chuang, I. Quantum Computation and Quantum Information. Am. J. Phys. 70, 558–559 (2002). 5. Levin, K. & Hulet, R. G. Chapter 3 - The Fermi Gases and Superfluids: Experiment and Theory. in Ultracold Bosonic and Fermionic Gases (eds. Levin, K., Fetter, A. L. & Stamper-Kurn, D. M.) vol. 5 69–94 (Elsevier, 2012). 6. Sobirey, L. et al. Observation of superfluidity in a strongly correlated two-dimensional Fermi gas. Science. 372, 844–846 (2021). 7. Bloch, I., Dalibard, J. & Nascimbène, S. Quantum simulations with ultracold quantum gases. Nat. Phys. 8, 267–276 (2012). 8. Labuhn, H. et al. Tunable two-dimensional arrays of single Rydberg atoms for realizing quantum Ising models. Nature 534, 667–670 (2016). 9. Monroe, C. et al. Programmable quantum simulations of spin systems with trapped ions. Rev. Mod. Phys. 93, 25001 (2021). 10. Zu, C. et al. Emergent hydrodynamics in a strongly interacting dipolar spin ensemble. Nature 597, 45–50 (2021). 11. Chomaz, L. et al. Dipolar physics: A review of experiments with magnetic quantum gases. arXiv (2022) doi:10.48550/ARXIV.2201.02672. 12. Christakis, L. et al. Probing site-resolved correlations in a spin system of ultracold molecules. arXiv (2022) doi:10.48550/ARXIV.2207.09328. 13. Yan, B. et al. Observation of dipolar spin-exchange interactions with lattice-confined polar molecules. Nature 501, 521–525 (2013). 14. Micheli, A., Brennen, G. K. & Zoller, P. A toolbox for lattice-spin models with polar molecules. Nat. Phys. 2, 341–347 (2006). 15. Barnett, R., Petrov, D., Lukin, M. & Demler, E. Quantum Magnetism with Multicomponent 17 Dipolar Molecules in an Optical Lattice. Phys. Rev. Lett. 96, 190401 (2006). 16. Hazzard, K. R. A., Manmana, S. R., Foss-Feig, M. & Rey, A. M. Far-from-Equilibrium Quantum Magnetism with Ultracold Polar Molecules. Phys. Rev. Lett. 110, 75301 (2013). 17. Mishra, T., Greschner, S. & Santos, L. Polar molecules in frustrated triangular ladders. Phys. Rev. A 91, 43614 (2015). 18. Bohn, J. L., Rey, A. M. & Ye, J. Cold molecules: Progress in quantum engineering of chemistry and quantum matter. Science 357, 1002–1010 (2017). 19. Gorshkov, A. V et al. Quantum magnetism with polar alkali-metal dimers. Phys. Rev. A 84, 33619 (2011). 20. Yao, N. Y., Zaletel, M. P., Stamper-Kurn, D. M. & Vishwanath, A. A quantum dipolar spin liquid. Nat. Phys. 14, 405–410 (2018). 21. Kwasigroch, M. P. & Cooper, N. R. Bose-Einstein condensation and many-body localization of rotational excitations of polar molecules following a microwave pulse. Phys. Rev. A 90, 21605 (2014). 22. Valtolina, G. et al. Dipolar evaporation of reactive molecules to below the Fermi temperature. Nature 588, 239–243 (2020). 23. Matsuda, K. et al. Resonant collisional shielding of reactive molecules using electric fields. Science 370, 1324–1327 (2020). 24. Li, J.-R. et al. Tuning of dipolar interactions and evaporative cooling in a three-dimensional molecular quantum gas. Nat. Phys. 17, 1144–1148 (2021). 25. Anderegg, L. et al. Observation of microwave shielding of ultracold molecules. Science 373, 779–782 (2021). 26. Schindewolf, A. et al. Evaporation of microwave-shielded polar molecules to quantum degeneracy. Nature 607, 677–681 (2022). 27. Bilitewski, T. et al. Dynamical Generation of Spin Squeezing in Ultracold Dipolar Molecules. Phys. Rev. Lett. 126, 113401 (2021). 28. Tobias, W. G. et al. Reactions between layer-resolved molecules mediated by dipolar spin exchange. Science 375, 1299–1303 (2022). 29. Kitagawa, M. & Ueda, M. Squeezed spin states. Phys. Rev. A 47, 5138–5143 (1993). 30. Leibfried, D. et al. Creation of a six-atom ‘Schrödinger cat’ state. Nature 438, 639–642 (2005). 31. Neyenhuis, B. et al. Anisotropic Polarizability of Ultracold Polar 40K87Rb Molecules. Phys. Rev. 18 Lett. 109, 230403 (2012). 32. Seebelberg, F. et al. Extending Rotational Coherence of Interacting Polar Molecules in a Spin-Decoupled Magic Trap. Phys. Rev. Lett. 121, 253401 (2018). 33. Burchesky, S. et al. Rotational Coherence Times of Polar Molecules in Optical Tweezers. Phys. Rev. Lett. 127, 123202 (2021). 34. Gullion, T., Baker, D. B. & Conradi, M. S. New, compensated Carr-Purcell sequences. J. Magn. Reson. 89, 479–484 (1990). 35. Geier, S. et al. Floquet Hamiltonian engineering of an isolated many-body spin system. Science 374, 1149–1152 (2021). 36. Ni, K.-K., Rosenband, T. & Grimes, D. D. Dipolar exchange quantum logic gate with polar molecules. Chem. Sci. 9, 6830–6838 (2018). 37. Eisert, J., Friesdorf, M. & Gogolin, C. Quantum many-body systems out of equilibrium. Nat. Phys. 11, 124–130 (2015). 38. Guardado-Sanchez, E. et al. Probing the Quench Dynamics of Antiferromagnetic Correlations in a 2D Quantum Ising Spin System. Phys. Rev. X 8, 21069 (2018). 39. Shenker, S. H. & Stanford, D. Black holes and the butterfly effect. J. High Energy Phys. 2014, 67 (2014). 40. Maldacena, J., Shenker, S. H. & Stanford, D. A bound on chaos. J. High Energy Phys. 2016, 106 (2016). 41. Braumüller, J. et al. Probing quantum information propagation with out-of-time-ordered correlators. Nat. Phys. 18, 172–178 (2022). 42. Green, A. M. et al. Experimental Measurement of Out-of-Time-Ordered Correlators at Finite Temperature. Phys. Rev. Lett. 128, 140601 (2022). 43. Davis, E., Bentsen, G. & Schleier-Smith, M. Approaching the Heisenberg Limit without Single-Particle Detection. Phys. Rev. Lett. 116, 53601 (2016). 44. Colombo, S. et al. Time-Reversal-Based Quantum Metrology with Many-Body Entangled States. Nat. Phys. 18, 925-930 (2022). 45. Gilmore, K. A. et al. Quantum-enhanced sensing of displacements and electric fields with two-dimensional trapped-ion crystals. Science 373, 673–678 (2021). 46. Bohn, J. L., Cavagnero, M. & Ticknor, C. Quasi-universal dipolar scattering in cold and ultracold gases. New J. Phys. 11, 55039 (2009). 19 47. Ticknor, C. Quasi-two-dimensional dipolar scattering. Phys. Rev. A 81, 42708 (2010). 48. Karman, T., Yan, Z. Z. & Zwierlein, M. Resonant and first-order dipolar interactions between ultracold 1S molecules in static and microwave electric fields. Phys. Rev. A 105, 13321 (2022). 49. Martin, M. J. et al. A Quantum Many-Body Spin System in an Optical Lattice Clock. Science 341, 632–636 (2013). 50. Bruun, G. M. & Taylor, E. Quantum Phases of a Two-Dimensional Dipolar Fermi Gas. Phys. Rev. Lett. 101, 245301 (2008). 51. Lee, H., Matveenko, S. I., Wang, D.-W. & Shlyapnikov, G. V. Fulde-Ferrell-Larkin-Ovchinnikov state in bilayer dipolar systems. Phys. Rev. A 96, 61602 (2017). 52. Peter, D., Müller, S., Wessel, S. & Büchler, H. P. Anomalous Behavior of Spin Systems with Dipolar Interactions. Phys. Rev. Lett. 109, 25303 (2012). 53. Koschorreck, M., Pertot, D., Vogt, E. & Köhl, M. Universal spin dynamics in two-dimensional Fermi gases. Nat. Phys. 9, 405–409 (2013). 54. Syzranov, S. V, Wall, M. L., Gurarie, V. & Rey, A. M. Spin–orbital dynamics in a system of polar molecules. Nat. Commun. 5, 5391 (2014). 20 Figure 1. A 2D itinerant spin system with polar molecules. a. Molecular spins are free to move within a 2D layer. Their dipolar interactions are tuned by a bias electric field E with configurable magnitude and orientation in the x-y plane. Yellow, blue, green, and grey curves illustrate interaction processes detailed in c. Red arrows represent dipolar elastic collisions. b. Energy diagram for the ground and first excited rotational states in which we encode the spin-1/2 degree of freedom. The transitions are driven by microwaves. c. Calculated dipolar interaction strength as a function of |E| for the spin manifold {|¯ñ, |1ñ}. The dipolar coupling strength is given in units of the permanent dipole moment of KRb dp = 0.574 Debye. The molecules interact via their induced dipole moments d¯ and d, as well as the transition dipole moment 𝑑↓↑. The black line shows the field dependence of 𝜒H. \n 21 Figure 2. Dynamical decoupling and tunable dipolar interactions between molecules. a. The standard Ramsey sequence consists of an initial pulse of area q, followed by an evolution time T, and finally a pulse of area p - q about an axis 𝑛n=\t𝑋b𝑐𝑜𝑠𝜙+\t𝑌b𝑠𝑖𝑛𝜙. We insert one or more XY8 sequences during T, denoted as XY8´M. One XY8 sequence (zoomed region) consists of 8 Rabi-p pulses spaced by time t, for a total free precession time of 8t (see Methods). b. Decay of Ramsey fringe contrast between |¯ñ and |1ñ without dynamical decoupling (grey circles) and with XY8´3 at a density of n = 0.14(2) ´ 107 cm-2 (orange squares). We use a low density to reduce interaction effects. Error bars are 1 standard deviation from bootstrapping (see Methods). c. Measured Ramsey fringes with dynamical decoupling for an average 2D density n = 1.4(1) ´ 107 cm-2 for initial population imbalance q = p/4 (blue diamonds), p/2 (black circles) and 3p/4 (red squares) for T = 1.2 ms at |E| = 0. Solid lines are fits to Acos[(p/180f) - Df] plus an offset, from which we extract the measured shift Df.d. Measured Df versus cosq for n =1.4(1) ´ 107 cm-2 (black circles) and 0.65(7) ´ 107 cm-2 (grey squares). The strength of the mean-field interaction Uc is extracted from the slope of the linear fit (solid lines). Error bars are 1 s.e.. e. Density dependence of Uc for the {|¯ñ, |1ñ} manifold (grey) and the {|¯ñ, |2ñ} manifold (green). Black and green solid lines are linear fits to the data. The slopes of the fits are direct measurements of c/ and c0 respectively. Error bars are 1 s.e. from linear fits. f. Dependence of c for the {|¯ñ, |1ñ} manifold on |E| at a = 0°. The solid line is a one parameter fit to 𝐴[(𝑑↓−𝑑↑)0−2𝑑↓↑𝑑↑↓] calculated for KRb at the experimental magnetic field and trapping conditions which includes modifications to the dipole moments from the mixing of hyperfine and rotational states at |E| below ~500 V/cm (shaded area). Grey line indicates zero. Error bars are 1 s.e. from linear fits. g. Angular dependence of c for the {|¯ñ, |1ñ} manifold at |E| = 1.02 kV/cm. Solid line is a one parameter fit to -A(3cos2a - 1). Grey line indicates zero. Error bars are 1 s.e. from linear fits. \n 22 Figure 3. Reversal of the spin dynamics. a. Measurement sequence. Molecules are initially prepared and allowed to evolve in a superposition of |¯ñ and |2ñ. A composite pulse R coherently transfers the population in |2ñ to |1ñ while preserving its relative phase to |¯ñ. Molecules then evolve in {|¯ñ ,|1ñ}. Inset: Composition of R. b. Measured time evolution of DF = Df(q = p/4) -Df(q = 3p/4) at n = 1.1(1) ´ 107 cm-2. The phase accumulation in stage I (green squares) and stage II (black circles) are plotted over time. For T <1.2 ms, we fix the duration of stage II to be 80 µs and scan the time of stage I. For T >1.2 ms, we fix the duration of stage I to be 1.2 ms and scan the time of stage II. The total time plotted excludes the widths of the microwave pulses. A piecewise linear fit to the time evolution of DF is shown as a solid line. The ratio between the two slopes is constrained to -2. Error bars are 1 s.e.. \n 23 \n Figure 4. Dipolar collisional decoherence. a. Decay of Ramsey contrast between |¯ñ and |1ñ for density n = 1.1(1) ´ 107 cm-2 (grey circles), 0.57(4) ´ 107 cm-2 (blue diamonds), 0.28(3) ´ 107 cm-2 (orange squares) at |E| = 0. Decoupling sequences used are XY8´3 for the grey and blue traces and XY8´7 for the orange trace. Error bars are 1 standard deviation from bootstrapping (see Methods). b. Contrast decay rate G as a function of n for |E| = 0, a = 0° (grey circles), and |E| = 1.02 kV/cm, a = 36° (purple squares). Error bars are 1 s.e. from exponential fits. \n" }, { "title": "1311.0965v1.Spin_accumulation_detection_of_FMR_driven_spin_pumping_in_silicon_based_metal_oxide_semiconductor_heterostructures.pdf", "content": "1 \n Spin accumulation detection of FMR driven spin pumping in silicon -based \nmetal -oxide -semiconductor hetero structures \n \nY. P u1, P. M. Odenthal2, R. Adur1, J. Beardsley1, A. G. Swartz2, D. V. Pelekhov1, R. K. \nKawakami2, J. Pelz1, P. C. Hammel1, E. Johnston -Halperin1 \n \n1Department of Physics, The Ohio State University , Columbus, Ohio 43210 \n2Department of Physics and Astronomy, University of California, Riverside, California 92521 \n \nThe use of the spin Hall effect and its inverse to electrically detect and manipulate dynamic \nspin currents generated via ferromagnetic resonance (FMR) driven spin pumping has \nenabled the investigation of these dynamically injected currents across a wide variety of \nferromagnetic materials. However, while this approach has proven to be an invaluable \ndiagnostic for exploring the spin pumping process it requires strong spin -orbit coupling , \nthus substantially limit ing the materials basis available for the detector/ch annel material \n(primarily Pt, W and Ta) . Here , we re port FMR driven spin pumping into a weak spin -\norbit channel through the measurement of a spin accumulation voltage in a Si-based metal -\noxide -semiconductor (MOS) heterostructure . This alternate experimental approach \nenables the investigation of dynamic spin pumping in a broad class of materials with weak \nspin-orbit coupling and long spin lifetime while providing additional information regarding \nthe phase evolution of the injected s pin ensemble via Hanle -based measurements of the \neffective spin lifetime . \n 2 \n The creation and manipulation of non -equilibrium spin populations in non -magnetic \nmaterials (NM) is one of the cornerstones of modern spintronics. These excitations have to date \nrelied primarily on charge based phenomena, either via direct electrical injection from a \nferromagnet (FM)1-7 or through the exploitation of the spin -orbit interaction8-10. Ferromagnetic \nresonance (FMR) driven spin pumping11-22 is an emerging method to dynam ically inject pure \nspin current into a NM with no need for an accompanying charge current, implying substantial \npotential impacts on low energy cost , high efficiency spintronics . However, while the creation of \nthese non -equilibrium spin currents does not require a charge current , previous studies of \ntransport -detected spin pumping do rely on a strong spin -orbit interaction in the NM to convert \nthe spin current into a charge current in the detector via the inverse spin -Hall effect (ISHE) 12-20. \nThis approach has proven to be an instrumental diagnostic , but it does carry with it several \nlimitations; specifically, the ISHE measures spin current not spin density , is only sensitive to a \nsingle component of the ful l spin vector and is only effective in materials with strong spin -orbit \ncoupling . Here we demonstrate an alternate detection geometry relying on the measurement of a \nspin accumulation voltage using a ferromagnetic electrode, similar to the three -terminal \ngeometry pioneered for electrically -driven spin injection4,23-28. This approach dramatically \nexpands the materials basis for FMR driven spin pumping, allows for the direct measurement of \nspin accumulation in the channel and enables the phase -sensitive meas urement of the injected \nspin population. \n Our study is performed in a silicon -based metal -oxide -semiconductor (MOS) structure \ncompatible with current semiconductor logic technologies. Using a Fe/MgO/p -Si tunnel diode \nwe achieve spin pumping into a semiconductor across an insulat ing dielectric. This approach \nallows voltage -based detection of the spin accumulation under the electrode4-6,23-29. Further, we 3 \n demonstrate sensitivity to the phase of the injected spin via the observation of Hanle dephasing \nin the presence of an out -of-plane magnetic field . These results establish a bridge between the \npure spin currents generated by FMR driven spin pumping and traditional charge -based spin \ninjection, laying the foundation for a new class of experimental probes and promising the \ndevelopment of novel spin -based devices compatible with current CMOS technologies. \n Tunne l diodes are fabricated from Fe(10nm)/ MgO( 1.3nm)/ Si(100) heterostructures \ngrown by mo lecular beam epitaxy (MBE). The p -type Si substrates are semiconductor on \ninsulator (SOI) wafers with a 3 μm thick Si device layer containing 5×1018 cm-3 boron dopants , \nproducing a room temperature resistivity of 2× 10-2 Ωcm. The device is patterned by conv entional \nphotolithography technique s into a Fe/MgO/Si tunnel contact of 500 μm × 500 μm lateral size, \nplaced 1 mm away from Au reference contacts for voltage measurements. Spin pumping and \nFMR measurements are performed in the center of a radio frequency (RF) microwave cavity \nwith f = 9.85GHz with a DC magnetic field , \nDCH , applied along the x-axis, as sketched in Fig. \n1a. On resonance , a pure sp in current is injected into the silicon channel via coupling between \nthe precessing magnetization of the ferromagnet, M, and the conduction electrons in the silicon. \nThis spin current induces an imbalance in the spin -resolved electrochemical potential and \nconsequent spin accumulation given by\n S , where \nand \n are the chemical \npotentials of up and down spin s, respectively. Using a standard electrical spin detection \ntechnique4-6,23-29 the spin accumulation can be detected electrically via the relationship : \n)1(2eP VS\nS\n \nwhere \nSV is the spin -resolved voltage between Fe and Si, \nP is the spin polarization of Fe , \n is \nthe spin detection efficiency , and \nS is assumed to be proportional to the component of the net 4 \n spin polarization parallel to M30. Figure 1c shows the magnetic field dependence of the FMR \nintensity (upper panel) and the spin accumulation induced voltage VS (lower panel), clearly \ndemonstrating spin accumulation at the ferromagnetic resonance . \n Figure 2a show s the RF power , PRF, dependence of the FMR intensity (upper panel) and \nSV\n (lower panel) on resonance ; the former is proportional to the square root of \nRFP and latter is \nlinear with\nRFP , consistent with ISHE detected spin pumping18-20. As shown in Fig.2b, \nSV is \nconstant when M reverses, consistent with our local detection geometry wherein the injected spin \nis always parallel to the magnetization of the FM electrode . Note that this is in contrast to the \nmagnetization dependence of ISHE detection, wherein the sign of the ISHE gives a measure of \nthe spin orientation relative to the detection electrode . As a result, o ur technique distinguishes the \nspin accumulation signal from artifacts due to magneto -transport or spin transport, such as the \nanomalous Hall effect, ISHE or spin Seebeck effect , which depend on the direction of M. In \naddition, the current -voltage characteristic ( I-V) of the tunnel contact is linear at room \ntemperature (see Supplementary Information ), implying at best a weak rectification of any RF -\ninduced pickup currents. This expectation is confirmed by the small offset (below 10 V) \nobserved in spin accumulation voltage measurement s. Our technique also rules out potential \nspurious signal s due to magneto -electric transport such as tunneling anisotropic magneto -\nresistance (TAMR) that would require a rectified bias voltage of at least mV scale to give the \nobserved V scale signal observed at resonance. \n In order to probe the dynamics of the observed spin accumulation, the applied magnetic \nfield is rotated towards the sample normal (within the xz-plane) by an angl e \n, introducing an \nout-of-plane component, \nzH. Due to the strong demagnetization field of our thin -film geometry \n(~2.2 T) the orientation of the magnetization lags the orientation of the applied field, remaining 5 \n almost entirely in -plane ( the maximum estimated deviation is 2°). As a result , the injected spins \n(parallel to M) precess due to the applied field. As the magnitude of \nzH increases this precession \nwill lead to a dephasing of the spin ensemble and consequent decrease in its net magnetization , \n(the Hanle effect3-6,23-29). Figure 3a shows the FMR spectrum for different angles \n ; the \nresonant field \nFMRH changes from 250G to 400G as \n changes from 0 to 40 degrees. This \nincrease is consistent with the fact that the in -plane component of H primarily determines the \nresonance condition, so as \n increases a larger total applied field is therefore required to drive \nFMR (see Supplementa ry Information ). Figure 3b shows \nSV vs. \nDCH over the same angular \nrange. The peak position of \nSV shifts in parallel with the FMR spectrum, but the peak value \ndecreases with increasing \nzH , as expected for Hanle -induced dephasing in an ensemble of \ninjected spins3-6,23-29. \nFor an isotropic ensemble of spins precessing in a uniform field perpendicular to M the \neffect of this dephasing on \nSV can be described by a simple Lorentzian function: \n)2()(1)(20\n SHS Vz x S S\n \nwhere \n is the Larmor frequency given by \n/z B effH g , \neffg is the effective Land é g-factor, \nB\n is the Bohr magneton ,\n is the Plank’s constant , and \n is the spin lifetime. In the more \ngeneral case that H is not perpendicular to M, as is the case here, then Eq. (2) should be replaced \nby the more general function: \n)3() (11\n2 22 2\n22\n0\n\n \n\ntotal totalz y\ntotalx\nx S S S V\n \nwhere \n),, (/ zyxi gHB effi i 23,27. If H is in the xz-plane, this reduces to 6 \n \n)4() (11sin cos22 2\n0\n\n\n \ntotalx S S S V \n This general behavior has been observed in previous studies of three terminal electrical \nspin injection4,23-28 (Fig. 4a). However, it has been widely reported in electrically detected spin \ninjection experiments that spatially varying local fields due to the magnetic electrodes, coupling \nto interfacial spin states and other non -idealities generate spin dynamics that are not well \ndescribed by this simple model. As a result \n is generally understo od to represent an effective \nspin lifetime, \neff , and while there are some initial efforts to more quantitatively account for the \nreal sample environment, such as the so -called “inver ted-Hanle” measurement23,26,27, a detailed \nmodel of these interactions is currently lacking. \nWe explore the functional dependence of the dephasing of our FMR driven spin current \nby plotting the peak spin accumulation voltage, \npeak\nSV , as a function of \nzH (Fig. 4b, solid \ncircles). The suppression of the spin accumulation at high magnetic fields seen in Fig. 3 is a clear \nindication of the dephasing of the injected spin ensemble; however, attempts to fit this behavior \nto Eq. (4) reveal that this simple, isotropic model fails to accurately reproduce our data (Fig. 4b, \nblack dashed lines). In particular, it is a feature of the isotropic model that for a magnetic field \nthat is not parallel to z that the spin polarization along M, and therefore the measured \naccumulation voltage, does not go to zero at high field even for infinite spin lifetime. This \ndiscrepancy likely arise s from contributions due to the various non -idealities discussed above ; in \nparticular, as we discuss below, we believe that the coupling to localized states and the impact of \nbulk spin diffusion may play a more central role in this experimental geometry. \nWe note that our data is well described by a simple Lorentzian (Eq. 2), though the \nrelationship between the effective spin lifetime extracted from this fit (0.6 ns) , which we label 7 \n \nFMR, to the \neff defined in Eq. (4) is not clear. For comparison, the intensity of the FMR signal \nis found to be constant to within roughly 10% (solid red triangles) , suggesting that the spin \ncurrent is roughly constant and indicating that FMR -driven heating19,20, if present, does not \ncontribute significantly to the field -induced suppression of \npeak\nSV seen i n Fig. 4b. \nA key advantage of our experimental geometry is that it allows direct comparison of this \ndephasing with the more traditional three -terminal electrical injectio n within the same device26. \nFigure 4b (open purple circles) shows the spin accumulation voltage measured in the three -\nterminal geometry as a function of a perpendicular applied field, \nzH . The dephasing in this \ngeometry is clearly much slower than in FMR driven s pin pumping. This observation is \nsupported by the Lorentzian fit to Eq. (2) indicated by the solid purple line, yielding an effective \nlifetime of 0.11 ns, consistent with previous reports by our group and others4,23-27. In considering \nthe origin of this discrepancy in observed lifetime a natural suspicion falls on the different \nexperimental methodologies . Specifically , for the spin pumping case the field is applied at an \nangle \n , resulting in both in -plane and perpendicular components to the field, while for the DC \ncurrent injection only the perpendicular component is present. \nThe consequence of this vector magnetic field is twofold: first, it will rotate the \nprecession axis of the injected spins away from the perpendicular case implicit in the simple \nHanle model as described above, and second, it will generate an “inver ted” Hanle effect that has \nbeen proposed to derive from the interplay between an in -plane applied magnetic fiel d and some \nfinite inhomogeneity in the local fields due to the magnetization of the electrode. In Fig. 4c w e \nexplore this behavior in a control sample wherein we perform both traditional Hanle (i.e. wherein \nthe only applied magnetic field is\nzH ) and rotating Hanle measurements (i.e. wherein the \nmagnetic field is rotated by an angle \n , yielding both in -plane and perpendicular components to 8 \n \nH), see Supplementary Information. The field values for the rotating Hanle experiment are \nchosen to correspond to the values o f \nzH from Fig. 4b. As expected, the traditional Hanle \nmeasurement again yields an effective lifetime of roughly 0.1 4 ns (open purple circles) . While \nthe rotating Hanle geometry does yield a slightly shorter lifetime of 0.09 ns (solid blue squares) , \nthis variation is too small (and in the wrong direction) to account for the longer effective lifetime \nobserved for FMR driven spin pumping. We therefore conclude that th e enhanced dephasing rate \nobserved in Fig. 4b indicates the FMR -driven and electrically -driven spin injection processes \ndiffer . \nWhile the origin of this discrepancy is still an active area of investigation, we note that \nthe observed \nFMR of 0.6 ns is consistent with previous measurements of the spin lifetime in the \nbulk silicon channel at these doping level4,23-25. Further, the requirement that there be no net \ncharge flow during FMR driven spin injection implies that any forward propagati ng tunneling \nprocess be balanced by an equal and opposite back tunneling process. As shown by the band \ndiagram of spin pumping in Fig. 4a this opens up a potential pathway for coupling from the bulk \nSi states into the intermediate states that dominate the three -terminal accumulation voltage26,28. \nThe band diagram of electrical spin injection (upper panel of Fig. 4a) shows that this process is \nstrongly suppressed in electrical spin injection due to the finite bias ( 5 mV in this case) present \nacross the tunnel junction . If we assume for the sake of argument that the spin pumping \nmeasurement is in fact sensitive to the bulk spin polarization in silicon, we can calculate the spin \ncurrent using the relation \nsf S SeJ ; according to Eq . (1) with \nP=0.4 for Fe and assuming \n\n=0.5, the spin current density is \nSJ ~ 2 × 105 Am-2. This spin current is roughly one order of \nmagnitude smaller than previous reports of spin pumping from conventional ferromagnets into \nmetals. 9 \n In summary, we demonstrate spin pumping into a semiconductor through an insulat ing \ndielectric . This approach allows observation of precession of the dynamically injected spins an d \ncharacteriz ation of the effective spin lifetime . Our results directly probe the coherence and phase \nof the dynamically injected spins and the spin manipulation of that spin ense mble via spin \nprecession, lay ing the foundation for novel spin pumping based spintronic applications . \n \nAcknowledgements \nThis work is supported by the Center for Emergent Materials at the Ohio State University, \na NSF Materials Research Science and Engineering Center (DMR -0820414) (YP, PO, JB, AS, \nRK, JP and EJH) and by the Department of Energy through grant DE -FG02 -03ER46054 (RA \nand PCH). Technical support is provided by the NanoSystems Laboratory at the Ohio State \nUniversity. The aut hors thank Andrew Berger and Steven Tjung for discussion s and assistance. \n \nReference and notes \n1. Žutić, I. , Fabian , J. & Das Sarma , S. Spintronics: Fundamentals and applications. Rev. Mod. \nPhys. 76, 323 (2004). \n2. Awschalom , D. D. & Flatté , M. E. Challenges for semiconductor spintronics. Nature Phys. \n3, 153 (2007). \n3. Appelbaum , I., Huang , B., & Monsma, D. J. Electronic measurement and control of spin \ntransport in silicon . Nature 447, 295 (2007). \n4. Dash, S. P. , Sharma, S., Patel, R. S. , de Jong, M. P. & Jansen, R. Electrical creation of spin \npolarization in silicon at room temperature . Nature 462, 491 (2009). 10 \n 5. Jedema, F. J., Heersche, H. B., Filip, A. T., Baselmans, J. J. A. & van Wees, B. J. Electrical \ndetection of spin precession in a metallic mesoscopic spin valve. Nature 416, 713 (2002). \n6. Lou, X. et al. Electrical detection of spin transport in lateral ferromagnet semiconductor \ndevices. Nature Phys. 3, 197 (2007). \n7. Jonker, B. T. , Kioscog lou, G., Hanbicki, A. T., Li, C. H., & Thompson, P. E. Electrical spin -\ninjection into silicon from a ferromagnetic metal /tunnel barrier contact . Nature Phys. 3, 542 \n(2007). \n8. Kato , Y. K., Myers , R. C., Gossard , A. C. & Awschalom , D. D. Observation of the Spin Hall \nEffect in Semiconductors . Science 306, 1910 (2004). \n9. Valenzuela , S. O. & Tinkham , M. Direct electronic measurement of the spin Hall effect . \nNature 442, 176 (2006). \n10. Liu, L. et al. Spin-Torque Switching with the Giant Spin Hall Effect of Tantalum. Science \n336, 555 (2012). \n11. Tserkovnyak Y., Brataas, A. & Bauer, G. E. W. Enhanced Gilbert damping in thin \nferromagnetic Films. Phys. Rev. Lett . 88, 117601 (2002). \n12. Saitoh, E., Ueda, M., Miyajima, H. & Tatara, G. Conversion of spin current into charge \ncurrent at room temperature: Inverse spin -Hall effect. Appl. Phys. Lett. 88, 182509 (2006). \n13. Tserkovnyak Y., Brataas, A. & Bauer, G. E.W. & Halperin, B. I. Nonlocal magnetization \ndynamics in ferromagnetic heterostructures. Rev. Mod. Phys. 77, 1375 (2005). \n14. Kajiwara , Y. et al . Transmission of electrical signals by spin -wave interco nversion in a \nmagnetic insulator . Nature 464, 262 (2010). \n15. Kurebayashi , H. et al . Controlled enhancement of spin -current emission by three -magnon \nsplitting . Nature Materials 10, 660 (2011). 11 \n 16. Sandweg, C. W . et al. Spin Pumping by Parametrically Excited Exchange Magnons . Phys. \nRev. Lett. 106, 216601 (2011). \n17. Czeschka , F. D. et al. Scaling behavior of the spin pumping effect in ferromagnet -platinum \nbilayers . Phys. Rev. Lett. 107, 046601 (2011). \n18. Ando , K. et al. Inverse spin -Hall effect induced by spin pumping in metallic system . J. Appl. \nPhys. 109, 103913 (2011). \n19. Ando , K. et al . Electrically tunable spin injector free from the impedance mismatch \nproblem . Nature Materials 10, 655 (2011). \n20. Ando , K. & Saitoh, E. Observation of the inverse spin Hall effect i n silicon. Nature Comm. \n3, 629 (2012 ). \n21. Costache, M. V. et al. Electrical detection of spin pumping due to the precessing \nmagnetization of a single ferromagnet. Phys. Rev. Lett. 97, 216603 (2006). \n22. Heinrich , B. et al . Spin pumping at the magnetic insulator (YIG)/normal metal (Au) \ninterfaces . Phys. Rev. Lett. 107, 066604 (2011) . \n23. Dash, S. P. et al. Spin precession and inverted Hanle effect in a semiconductor near a finite -\nroughness ferromagnetic interface . Phys. Rev. B 84, 054410 (2011). \n24. Li, C. H., van‘t Erve , O. & Jonker , B. T. Electrical injection and detection of spin \naccumulation in silicon at 500K with magnetic metal/silicon dioxide contacts . Nature \nComm. 2, 245 (2011 ). \n25. Gray , N. W. & Tiwaria , A. Room temperature electrical injection and detection of spin \npolarized carriers in silicon using MgO tunnel barrier . Appl. Phys. Lett. 98, 102112 (2011 ). \n26. Pu, Y . et al. Correlation of electrical spin injection and non -linear charge -transport in \nFe/MgO/Si . Appl. Phys. Lett. 103, 012402 (2013 ). 12 \n 27. Jeon, K. et al. Electrical investigation of the oblique Hanle effect in ferromagnet / oxide / \nsemiconductor c ontacts . arXiv 1211.3486 (2013). \n28. Tran, M. et al. Enhancement of the spin accumulation at the i nterface between a spin-\npolarized tunnel junction and a semiconductor . Phys. Rev. Lett. 102, 036601 (2009 ). \n29. Sasaki, T., Oikawa, T., Suzuki, T., Shiraishi, M., Suzuki, Y. & Noguchi, K. Comparison of \nspin signals in silicon between nonlocal four -terminal and three -terminal methods. Appl. \nPhys. Lett. 98, 012508 (2011). \n30. As discussed in Ref. 26, Eq. (1) is derived assuming a linear tunneling model, and might \nunderestimate the actual value of \nS at higher bias if the current depends super -linearly on \napplied bias. \n 13 \n Figure legends \n \nFigure 1 | Experimental setup \n(a) Schematic of experimental setup. ( b) Diagram of spin accumulation and spin -resolved \nvoltage \nSV. (c) FMR intensity (upper panel; arrows indicate state of the Fe magnetiza tion) and \nspin-resolved voltage \nSV (lower panel) as a function of \nDCH . \n \nFigure 2 | RF power - and magnetic field - dependence \n(a) FMR intensity (upper panel) and spin-resolved voltage (lower panel) as a function of RF \npower; ( b) Solid symbols: \nSV vs. \nFMR DC H H when \nDCH is parallel or anti -parallel with the x-\naxis; open symbols indicate the voltage between two Au/Si reference contacts; all data is \nmeasured under the same experimental conditions. A background offset of ~2 V has been \nsubtracted from all data. \n \nFigure 3 | Experiments with increasing Hz \n(a) FMR intensity spectra at various magnetic field orientations \n as described in the text; ( b) \nSV\n vs. \nDCH measured at the same set of magnetic field orientations. The shift in FMR center \nfrequency tracks the expected magnetization anisotropy of the Fe thin film, see text. \n \nFigure 4 | Hanle effect measurements 14 \n (a) Schematics of experimental setup and band diagram for three -terminal electrical spin \ninjection (upper panel) and spin pumping (lower panel); ( b) Hanle effect as a function of \nzH for \nthree -terminal (open circles) and spin pumping (solid circles), solid lines are Lorentzian fits \nyielding 0.11 ns and 0.6 ns, respectively; solid triangles are FMR absorption as a function of H z, \nthe red dashed line is a guide to the eye and the scale bar represents 10% variation; Black dashed \nlines are simulated using Eq. (4) with \n , \nxH = 248G ( dash dot ) and 310G ( dash), \nrespectively , see Supplementary Information . (c) Hanle effect measured in a control sample by \nthe three -terminal method, open circles are measured with magnetic field applied out of plane, \nsolid squares are obtained using same magnetic field configuration as for FMR driven spin \npumping; lines are Lorentzian fits yielding 0.14 ns and 0.09 ns, respectively. 15 \n Figures \n \n \n \nFigure 1 Y. Pu et al. \n16 \n \n \n \n \nFigure 2 Y. Pu et al. \n \n17 \n \n \n \n \nFigure 3 Y. Pu et al. \n \n18 \n \n \n \n \nFigure 4 Y. Pu et al. \n19 \n Supplementary Information \n \nSpin accumulation detection of FMR driven spin pumping in silicon -based \nmetal -oxide -semiconductor heterostructures \n \nY. Pu1, P. M. Odenthal2, R. Adur1, J. Beardsley1, A. G. Swartz2, D. V. Pelekhov1, R. K. \nKawakami2, J. Pelz1, P. C. Hammel1, E. Johnston -Halperin1 \n \n1Department of Physics, The Ohio State University, Columbus, Ohio 43210 \n2Department of Physics and Astronomy, University of California, Riverside, California 92521 \n \nA. Linear I -V of Fe/MgO/Si contact at room temperature \n The I-V characteristic of the Fe/MgO/p -Si contact is linear at room temperature , as \nshown in Fig. S1, indicating that in this regime the contact resistance is domin ated by the MgO \ninsulating barrier and contribution from the Schottky barrier is negligible. The linear fit gives \n40.9 resistance with about 25m uncertainty. \n \nB. Impact of interface roughness and in -plane external magnetic field \n In contrast to traditional three -terminal spin accumulation measurements, for the FMR \ndriven measurements described in Figs. 3 and 4 it is necessary to apply an external magnetic \nfield both parallel and perpendicular to the magnetization. As described in the main text, the \nparallel component of the field is necessary to satisfy the conditions of magnetic resonance and 20 \n the perpendicular component contributes to the decay of spin accumulation, allowing for a \nHanle -style measurement of the effective spin life time. However, in consulting the literature23,27 \nit becomes evident that this geometry potentially raises an additional concern regarding the \ninterpretation of this data. Specifically, the in plane component of the magnetic field will itself \ninduce an “inv erted -Hanle” effect wherein the measured spin accumulation rises with the \nmagnitude of the parallel component of the magnetic field (Fig. S2a). The accepted interpretation \nof this effect is an annealing of fluctuations in the magnetization induced by surfa ce roughness of \nthe magnetic layer in large external fields23,27. \n We measure the spin accumulation on a control sample via 3T electrical spin injection, as \nshown in Fig. S2a, where magnetic field is applied in xz -plane with orientation ranging from in -\nplane ( H//x, 0 degree) to out of plane (H//z, 90 degree). Using the 3T data of \n) (FMR S H HV , \nwhere \nFMRH for given magnetic field orientation is obtained by spin pumping experiment as \nshown in Figure 3, we can first directly compare electrical spin injection and spin pumping under \nthe same experimental configuration, as discussed in main text. \nTo get a quantitative understanding of the angular dependence shown in Fig. S2a, one can \nstart with a general formula: \nSSωSSDt\n (S1) \nwhere \n is the Larmor frequency , \nD is the spin diffusion constant and \n is the spin lifetime. In \nthe 3T geometry the impact of spin diffusion is usually believed to be negligible23,27, from this \none can obtain an analytical solution under arbitrary applied magnetic field: 21 \n \n\n\n 2 22 2\n22\n0) (11\n \n\ntotal totalz y\ntotalx\nxS S (S2) \nwhere \n),, (/ zyxi gHB effi i , representing each component of the applied magnetic field. \nFigure S2b shows the simulation according to Equation (S2) with \n = 0.14ns as obtained by \nLorentzian fit. Clearly the model fails to explain the data shown in Fig. S2a, especially the \nobservation that at certain orientations the measured spin accumulation rises with the applied \nmagnetic field increasing. \n As pointed out by previous studies23,27 stray magnetic fields from the injector due to \ninterface roughness should strongly impact on the Hanle -style measurements. The total magnetic \nfield should be taken as\n),, ( zyxi H H Hms\niext\ni i , which re presents the contribution from \nexternal and magnetic -stray fields. The stray field strength is taken to have a spatial variation \n)/2cos()0( )( x Hx Hms\nims\ni \n, where \n is the typical length scale (~20nm) of the surface \nroughness23,27. Assuming the spin diffusion length is much longer than \n we average the total \nmagnetic field over a full period of \n , i.e. \n2 2 2) () (ms\niext\ni i H H H , we therefore have a formula: \n\n\n\n2 2 2 22 2\n22 2\n0)/ (11 ) () () () (\n total B totalext\nzms\ntotalext\nxms\nx\nxH g HH H\nHH HS S\n (S3) \nwhere \nms\nxH and\nmsH represent the averaged stray field parallel or perpendicular to the injected \nspins, respectively. Figure S2c shows the simulation according to Equation (S3) with parameter s \n\n= 0.9ns, \nms\nxH 270G and \nmsH 440G. The simulation qualitatively agrees with the \nexperiment, but shows some systematic deviations especially in the low -field regime. 22 \n Although the Equation (S1) is generally accepted, and in principle rigorous analysis can \nbe done with spin precessi on, spin diffusion and spin flip involved, a well -established approach \nto determine the intrinsic spin lifetime using the local spin detection geometry is still lacking. A \nprecise determination of the intrinsic spin lifetime in our sample is beyond the sco pe of this \nreport; we treat the lifetime obtained by the simple Lorentzian fit as an effective spin lifetime or \nspin dephasing time, which represents the decay rate of average spin accumulation under applied \nperpendicular magnetic field. \n \nC. Simulation on the Hanle effect under FMR condition \nAs indicated by the FMR spectrum, at FMR there are varying x- and z- components of the \napplied magnetic field with different field orientation \n . The table below is a summary: \n (deg.) 0 10 20 25 30 40 60 \nHx (G) 250 276 287 295 310 302 248 \nHz (G) 0 49 105 138 179 253 430 \n \nAs shown in the table, \nzH increases monotonically with \n and \nxH is in range of 248 – 310 G \n(roughly constant to maintain the conditions for magnetic resonance), both should impact on the \nHanle effect. The black dashed curves shown in Fig. 4(b) are simulated using Eq. (4) with \n\nand \nxH = 248G, 310G respectively. In the situation that Eq. (4) is valid, finite values of the spin \nlifetime should give a weaker H z-dependence than the simulation curve. 23 \n \n \nFigure S1: Plot of current vs. voltage of the Fe/MgO/p -Si contact at room temperature, symbol s \nare data and the solid line is a linear fit. \n \n \n \n \n24 \n \n \nFigure S 2: (a) Spin accumulation from 3T electrical spin injection, magnetic field is applied in \ndifferent orientations, ranging from in -plane ( H//x, 0 degree, top curve) to out of plane ( H//z, 90 \ndegree, bottom curve); (b) Simu lations according to Equation (S2) with \n= 0.14ns, assuming no \nstray field; ( c) Simulations using Equation (S3 ) with parameters \n= 0.9ns, \nms\nxH 270G and \nmsH\n440G . \n \n \n \n" }, { "title": "1704.02707v1.Majorana_dynamical_mean_field_study_of_spin_dynamics_at_finite_temperatures_in_the_honeycomb_Kitaev_model.pdf", "content": "APS/123-QED\nMajorana dynamical mean-\feld study of spin dynamics at \fnite temperatures in the\nhoneycomb Kitaev model\nJunki Yoshitake,1Joji Nasu,2Yasuyuki Kato,1and Yukitoshi Motome1\n1Department of Applied Physics, University of Tokyo, Bunkyo, Tokyo 113-8656, Japan\n2Department of Physics, Tokyo Institute of Technology, Meguro, Tokyo 152-8551, Japan\n(Dated: April 11, 2017)\nA prominent feature of quantum spin liquids is fractionalization of the spin degree of freedom.\nFractionalized excitations have their own dynamics in di\u000berent energy scales, and hence, a\u000bect\n\fnite-temperature ( T) properties in a peculiar manner even in the paramagnetic state harboring\nthe quantum spin liquid state. We here present a comprehensive theoretical study of the spin\ndynamics in a wide Trange for the Kitaev model on a honeycomb lattice, whose ground state is\nsuch a quantum spin liquid. In this model, the fractionalization occurs to break up quantum spins\ninto itinerant matter fermions and localized gauge \ruxes, which results in two crossovers at very\ndi\u000berentTscales. Extending the previous study for the isotropic coupling case [J. Yoshitake, J. Nasu,\nand Y. Motome, Phys. Rev. Lett. 117, 157203 (2016)], we calculate the dynamical spin structure\nfactorS(q;!), the NMR relaxation rate 1 =T1, and the magnetic susceptibility \u001fwhile changing\nthe anisotropy in the exchange coupling constants, by using the dynamical mean-\feld theory based\non a Majorana fermion representation. We describe the details of the methodology including the\ncontinuous-time quantum Monte Carlo method for computing dynamical spin correlations and the\nmaximum entropy method for analytic continuation. We con\frm that the combined method provides\naccurate results in a wide Trange including the region where the spins are fractionalized. We\n\fnd that also in the anisotropic cases the system exhibits peculiar behaviors below the high- T\ncrossover whose temperature is comparable to the average of the exchange constants: S(q;!) shows\nan inelastic response at the energy scale of the averaged exchange constant, 1 =T1continues to\ngrow even though the equal-time spin correlations are saturated and almost Tindependent, and\n\u001fdeviates from the Curie-Weiss behavior. In particular, when the exchange interaction in one\ndirection is stronger than the other two, the dynamical quantities exhibit qualitatively di\u000berent T\ndependences from the isotropic case at low T, re\recting the opposite parity between the \rux-free\nground state and the \rux-excited state, and a larger energy cost for \ripping a spin in the strong\ninteraction direction. On the other hand, when the exchange anisotropy is in the opposite way,\nthe results are qualitatively similar to those in the isotropic case. All these behaviors manifest\nthe spin fractionalization in the paramagnetic region. Among them, the dichotomy between the\nstatic and dynamical spin correlations is unusual behavior hardly seen in conventional magnets.\nWe discuss the relation between the dichotomy and the spatial con\fguration of gauge \ruxes. Our\nresults could stimulate further experimental and theoretical analyses of candidate materials for the\nKitaev quantum spin liquids.\nI. INTRODUCTION\nQuantum many-body systems show various intriguing\nphenomena which cannot be understood as an assembly\nof independent particles. One of such phenomena is frac-\ntionalization, in which the fundamental degree of freedom\nin the system is fractionalized into several quasiparticles.\nA well-known example of such fractionalization is the\nfractional quantum Hall e\u000bect, in which the Hall conduc-\ntance shows plateaus at fractional values of e2=h(eis the\nelementary charge and his the Planck constant) [1, 2].\nIn this case, the quasiparticles carry a fractional value of\nthe elementary charge, as a collective excitation of the\nelementary particle, electron. This is fractionalization of\ncharge degree of freedom. On the other hand, another\ndegree of freedom of electrons, spin, can also be fraction-\nalized. Such a peculiar phenomenon has been argued for\nquantum many-body states in insulating magnets, e.g., a\nquantum spin liquid (QSL) state.\nQSLs are the magnetic states which preserve all the\nsymmetries in the high-temperature( T) paramagnet evenin the ground state and evade a description by conven-\ntional local order parameters. A typical example of QSLs\nis the resonating valence bond (RVB) state, proposed by\nP. W. Anderson [3]. The RVB state is a superposition\nof valence bond states (direct products of spin singlet\ndimers), which does not break either time reversal or\ntranslational symmetry. In the RVB state, the spin de-\ngree of freedom is fractionalized: the system exhibits two\ndi\u000berent types of elementary excitations called spinon\nand vison [4, 5]. Spinon is a particlelike excitation car-\nrying no charge but spin S= 1=2. Meanwhile, vison\nis a topological excitation characterized by the parity\nof crossing singlet pairs with its trace. Another exam-\nple of QSLs is found in quantum spin ice systems, in\nwhich peculiar excitations are assumed to be magnetic\nmonopoles, electric gauge charges, and arti\fcial photons\nresulting from fractionalization of the spin degree of free-\ndom [6, 7].\nAmong theoretical models for QSLs, the Kitaev model\nhas attracted growing interest, as it realizes the frac-\ntionalization of quantum spins in a canonical form [8].arXiv:1704.02707v1 [cond-mat.str-el] 10 Apr 20172\nThe Kitaev model is a localized spin S= 1=2 model de-\n\fned on a two-dimensional honeycomb lattice with bond-\ndependent anisotropic interactions (see Sec. II A). In this\nmodel, the ground state is exactly obtained as a QSL,\nin which quantum spins S= 1=2 are fractionalized into\nitinerant Majorana fermions and localized gauge \ruxes.\nThe fractionalization a\u000bects the thermal and dynami-\ncal properties in this model. For instance, the di\u000berent\nenergy scales between the fractionalized excitations ap-\npear as two crossovers at largely di\u000berent Tscales; in\neach crossover, itinerant Majorana fermions and local-\nized gauge \ruxes release their entropy, a half of log 2\nper site [9, 10]. Also in the ground state, the dynamical\nspin structure factor S(q;!) shows a gap due to the \rux\nexcitation and strong incoherent spectra from the com-\nposite excitations between itinerant Majorana fermions\nand localized gauge \ruxes [11]. Such incoherent spectra\nwere indeed observed in recent inelestic neutron scatter-\ning experiments for a candidate for the Kitaev QSL, \u000b-\nRuCl 3[12, 13]. The magnetic Raman scattering spectra\nalso shows a broad continuum dominated by the itinerant\nMajorana fermions, in marked contrast to conventional\ninsulating magnets [14]. Such a broad continuum was ex-\nperimentally observed also in \u000b-RuCl 3[15]. Furthermore,\ntheTdependence of the incoherent response was theoret-\nically analyzed and identi\fed as the fermionic excitations\nemergent from the spin fractionalization [16, 17].\nIn the previous study, the authors have studied dy-\nnamical properties of the Kitaev model at \fnite Tby a\nnewly developed numerical technique, the Majorana dy-\nnamical mean-\feld method [18]. Indications of the spin\nfractionalization were identi\fed in the Tdependences of\nS(q;!), the relaxation rate in the nuclear magnetic res-\nonance (NMR), 1 =T1, and the magnetic susceptibility \u001f.\nIn the previous study, however, the results were limited\nto the case with the isotropic exchange constants, despite\nthe anisotropy existing in the Kitaev candidate materi-\nals [19, 20]. In the present paper, to complete the anal-\nysis, we present the numerical results of the dynamical\nquantities for anisotropic cases. We also provide the com-\nprehensive description of the theoretical method, includ-\ning the details of the cluster dynamical mean-\feld theory\n(CDMFT), the continuous-time quantum Monte Carlo\n(CTQMC) as a solver of the impurity problem to calcu-\nlate the dynamical spin correlations, and the maximum\nentropy method (MEM) for the analytic continuation.\nWe discuss a prominent feature proximate to the QSL,\ni.e., the dichotomy between static and dynamical spin\ncorrelations, from the viewpoint of the fractionalization\nof spins.\nThe paper is organized as follows. In Sec. II, after in-\ntroducing the Kitaev model and its Majorana fermion\nrepresentation, we describe the details of the CDMFT,\nCTQMC, and MEM. In Sec. III, we show the numeri-\ncal results for S(q;!), 1=T1, and\u001fwhile changing the\nanisotropy in the exchange constants. In Sec. IV, we\ndiscuss the dichotomy between the static and dynamical\nspin correlations by comparing the Tdependences for the\nK2 ΓM1\nM2K1(a) (b)FIG. 1. (a) Schematic picture of the Kitaev model on the\nhoneycomb lattice. The blue, green, and red bonds represent\nthep=x;y, andzbonds in Eq. (1), respectively. The dashed\noval represents the 26-site cluster used in the CDMFT calcu-\nlations. (b) The \frst Brillouin zone (black hexagon) and the\nsymmetric lines (red lines) used in Figs. 3 and 4.\nuniform and random \rux con\fgurations. The cluster-size\ndependence in the CDMFT is examined in Appendix A.\nThe accuracy of MEM is also examined in Appendix B in\nthe one-dimensional limit where the dynamical properties\ncan be calculated without analytic continuation. We also\nshow theTand!dependence of spin correlations and the\nTdependence of the Korringa ratio in Appendix C and\nD, respectively.\nII. MODEL AND METHOD\nIn this section, we describe the details of the meth-\nods used in the present study, the Majorana CDMFT\nand CTQMC methods. After introducing the Majorana\nfermion representation of the Kitaev model in Sec. II A,\nwe describe the framework of the Majorana CDMFT in\nSec. II B, in which the impurity problem is solved exactly.\nIn Sec. II C, we introduce the CTQMC method which is\napplied to the converged solutions obtained by the Majo-\nrana CDMFT for calculating dynamical spin correlations.\nWe also touch on the MEM used for obtaining the dy-\nnamical spin correlations as functions of real frequency\nfrom those of imaginary time in Sec. II D.\nA. Kitaev model and the Majorana fermion\nrepresentation\nWe consider the Kitaev model on a honeycomb lattice,\nwhose Hamiltonian is given by [8]\nH=\u0000X\npJpX\nhj;j0ipSp\njSp\nj0; (1)\nwherep=x,y, andz, and the sum of hj;j0ipis taken\nfor the nearest-neighbor (NN) sites on three inequiv-\nalent bonds of the honeycomb lattice, as indicated in\nFig. 1(a);Sp\njis thepcomponent of the S= 1=2 spin\nat sitej. Hereafter, we denote the average of JpasJ3\nand set the energy scale asP\npjJpj= 3, i.e.,J= 1, and\nparametrize the anisotropy of the exchange coupling con-\nstants asJx=Jy=\u0006\u000bandJz=\u0006(3\u00002\u000b), where +\nand\u0000correspond to the ferromagnetic (FM) and anti-\nferromagnetic (AFM) cases, respectively. We note that\nthe FM and AFM cases are connected through unitary\ntransformations [8].\nAs shown by Kitaev [8], the model is soluble and the\nexact ground state is obtained as a QSL. The spin corre-\nlations are extremely short-ranged: hSp\njSp\nj0iare nonzero\nonly for the NN sites j;j0on thepbonds as well as the\nsame sitej=j0[23]. Hereafter, we denote the NN corre-\nlations ashSp\njSp\nj0iNN. There are two types of QSL phases\ndepending on the anisotropy in Jp: one is a gapless QSL\nrealized in the region with 0 :75\u0014\u000b\u00141:5 including the\nisotropic point \u000b= 1 (Jx=Jy=Jz=\u00061), while the\nother is gapful for 0 \u0014\u000b < 0:75. The ground state\nhas nontrivial fourfold degeneracy in the thermodynamic\nlimit [24].\nThe exact solution for the ground state was originally\nobtained by introducing four types of Majorana fermions\nfor eachS= 1=2 spin [8]. In this method, the Hilbert\nspace in the original spin representation, 2N, is extended\nto 4Nin the Majorana fermion representation ( Nis the\nnumber of spins). Thus, to calculate physical quantities,\nsuch as spin correlations, it is necessary to make a pro-\njection from the extended Hilbert space to the original\none.\nSoon later, however, another way of solving the model\nwas introduced by using only two types of Majorana\nfermions [25{27], in which the projection is avoided as\nthe Hilbert space is not extended. In this method, the\nspin operators are written by spinless fermions by ap-\nplying the Jordan-Wigner transformation to the one-\ndimensional chains composed of two types of bonds, say,\ntheJxandJybonds. Then, by introducing two Ma-\njorana fermions cjand \u0016cjfor the spinless fermions, the\nHamiltonian in Eq. (1) is rewritten as\nH=iJx\n4X\n(j;j0)xcjcj0\u0000iJy\n4X\n(j;j0)ycjcj0\u0000iJz\n4X\n(j;j0)z\u0011rcjcj0;\n(2)\nwhere the sum over ( j;j0)pis taken for the NN sites on a\npbond withj T Hfor both uniform and random f\u0011gsim-\nilar to the result by the CDMFT+CTQMC method in\nRef. [18], it shows di\u000berent behavior below THbetween\nthe two cases, as shown in Fig. 12(d). For the case with\nuniformf\u0011g, 1=T1decreases to zero after showing a small\nhump. The suppression at low Tre\rects the \rux gap\n\u0001'0:065Jin the \rux-free state [8, 11]. On the other\nhand, for the case with random f\u0011g, 1=T1monotonically\nincreases while decreasing Tin the calculated Trange.\nSimilarTdependences of 1 =Tp\n1are obtained for 1 =Tx\n1at\n\u000b= 0:8 and 1=Tx;z\n1at\u000b= 1:2, as shown in Figs. 12(e)\nand 12(f), respectively. We note that 1 =Tzfor\u000b= 0:8\nbehaves di\u000berently; we will comment on this point in the\nend of this section.\nThe results clearly indicate that the peculiar Tde-\npendences of 1 =T1found in the CDMFT+CTQMC re-\nsults are closely related with \ructuations of the gauge\n\ruxesf\u0011gcomposed of localized Majorana fermions f\u0016cg\nemergent from the spin fractionalization. As seen in\nthe equal-time spin correlations shown in Figs. 12(a)-\n12(c), itinerant matter fermions develop their kinetic en-\nergy to the saturation at T\u0018TH(the equal-time spin\ncorrelations correspond to the kinetic energy of mat-\nter fermions). Due to the fractionalization, however,\nthe localized gauge \ruxes are still disordered even below\nTH[10], which results in the enhancement of 1 =T1, as in-\ndicated in Figs. 12(d)-12(f). When approaching TL,f\u0011g\nare aligned in a coherent manner [10], and hence, 1 =T1is\nrapidly suppressed at T\u0018TL. Thus, the Tdependence\nof 1=T1is qualitatively explained by the crossover from\nthat for the random f\u0011gto the fully-aligned f\u0011gwhile\ndecreasing T. The crossover occurs well below THand\nclose toTL. Of course, as the original quantum spin is\na composite of itinerant matter fermions and localized\ngauge \ruxes, the spin-\rip dynamics is a composite exci-\ntation. Nevertheless, our results indicate that the pecu-\nliarTdependence of the NMR relaxation rate as well as15\nT T T(a) (b) (c)\n(d) (e) (f)z x\nuniform\nrandom\nCDMFT\n10-210-110010-210-11000.00.20.40.60.81.0\n10-310-210-1100101\n10-210-1100uniform\nrandom\nCDMFTz x\nuniform\nrandom\nCDMFT\nuniform\nrandom\nCDMFT+CTQMCuniform\nrandom\nCDMFT+CTQMCz x\nuniform\nrandom\nCDMFT+CTQMCz x\nFIG. 12. (a)(b)(c) 4 hSp\njSp\nj0iNNfor the FM case and (d)(e)(f) the onsite components of 1 =Tp\n1(p=z;x) calculated by setting\nall\u0011= 1 (uniform) and all \u0011being random (random) in the CDMFT calculations: (a)(d) \u000b= 1:0, (b)(e)\u000b= 0:8, and (c)(f)\n\u000b= 1:2. In (a) and (d), hSp\njSp\nj0iNNand 1=Tp\n1are equivalent for p=x;z. For comparison, we plot the data in Figs. 2 (CDMFT)\nand 9 (CDMFT+CTQMC). The vertical dotted lines represent TLandTHfor each\u000b.\nthe magnetic susceptibility is dominated by the emergent\ngauge \ruxes from the fractionalization.\nAs noted above, 1 =Tz\n1for\u000b= 0:8 behaves di\u000berently\nfrom others: 1 =Tz\n1for the randomf\u0011gis smaller than\nthat for the uniform f\u0011gat lowT, as shown in Fig. 12(e).\nThis is presumably because of the peculiar Tdependence\nof the density of states (DOS) for the itinerant matter\nfermions at \u000b= 0:8. In the gapless QSL region for\n0:75\u0014\u000b\u00141:5 but close to the gapless-gapful boundary\nat\u000b= 0:75, the DOS opens a gap as f\u0011gare thermally\ndisordered by raising T[10]. Thus, the DOS for matter\nfermions is gapless for the uniform f\u0011g, while gapped for\nthe randomf\u0011g. As spin excitations by Sx\njandSy\njare\ncomposite excitations of both itinerant matter fermions\nand localized gauge \ruxes, the gap in the DOS for matter\nfermions suppresses 1 =Tz\n1for the random case compared\nto the uniform one. Since f\u0011gare aligned uniformly be-\nlowTL, we expect that 1 =Tz\n1shows an abrupt increase\nwhile decreasing TthroughTL. This indicates that while\na rapid change of 1 =T1when approaching TLis yielded\nby the coherent alignment of f\u0011g, either increase or de-\ncrease of 1=T1atTLmay be a\u000bected by the itinerant\nmatter fermions.\nV. SUMMARY\nTo summarize, we have presented numerical results for\nspin dynamics of the Kitaev model with the anisotropyin the bond-dependent coupling constants. We calculated\nthe experimentally-measurable quantities, the dynamical\nspin structure factor S(q;!), the NMR relaxation rate\n1=T1, and the magnetic susceptibility \u001f, in the wide T\nrange including the peculiar paramagnetic region where\nquantum spins are fractionalized. The results have been\nobtained by the Majorana CDMFT+CTQMC method,\nwhich were developed by the authors previously [18]; we\ngave detailed descriptions of the method, including the\nMEM for analytical continuation. We also con\frmed the\nMajorana CDMFT is precise enough in the range of T\nand anisotropy that we investigated in the present study.\nWe found that the Kitaev model exhibits unconven-\ntional behaviors in spin dynamics in the \fnite- Tparam-\nagnetic state in proximity to the QSL ground state. The\nprominent feature is the dichotomy between static and\ndynamical spin correlations as a consequence of the spin\nfractionalization. The dichotomy appears clearly in the\nincrease of 1 =T1belowTHwhere the fractionalization sets\nin, despite the saturation of static correlations. Similar\nbehavior was also seen in the isotropic case in the previ-\nous study [18]. Our results suggest that the dichotomy\nis found universally in the fractionalized paramagnetic\nregion irrespective of the anisotropy in the system.\nOn the other hand, we also clari\fed interesting behav-\niors that depend on the anisotropy at low T. When one\nof the three bond-dependent interactions is stronger than\nthe other two, the spin dynamics shows peculiar Tand\nenergy dependences distinct from those in the isotropic16\ncoupling case as follows. As lowering T,S(q;!) develops\na\u000e-function like peak, which is well separated from the\nincoherent continuum. 1 =T1monotonically decreases in\nthe spin component for the stronger bond. \u001fincreases\nand saturates to a nonzero value for the spin component\nfor the weaker bonds, while it shows hump and then de-\ncreases for the stronger-bond component in the antifer-\nromagnetic case. We also showed that the peculiar Tde-\npendences of \u001fare qualitatively explained by the two-site\ndimer model. In contrast, when the anisotropy is oppo-\nsite, i.e., when the two types of bonds become stronger,\nthe results are qualitatively unchanged from those for the\nisotropic case, while the e\u000bect of anisotropy is obvious in\ntheqdependence in S(q;!) and the di\u000berent components\nin 1=T1and\u001f.\nOur results will stimulate further experimental and\ntheoretical analyses of candidate materials for the Ki-\ntaev QSLs. As most of the materials are assumed to be\nanisotropic in the exchange constants [19{22], our results\nwill be helpful for understanding of unusual behaviors\nin the real compounds. We emphasize that our numer-\nical data obtained by the Majorana CDMFT+CTQMC\nmethod are quantitatively reliable in the calculated para-\nmagnetic regime, as the cluster approximation and the\nanalytic continuation are both well controlled. Although\nthere are residual interactions in addition to the Kitaev-\ntype ones in real materials, our results provide good ref-\nerences in the limit of the pure Kitaev model for inter-\npreting the role of the additional interactions.\nWhile we have calculated dynamical quantities of the\nKitaev model in the wide Trange, the calculations were\nlimited above TLdue to the phase transition which is\nartifact of the mean-\feld nature of CDMFT. It is nec-\nessary to develop more sophisticated method to study\nthe dynamical properties below TL. The low-Tspin dy-\nnamics will be interesting, in particular, for extensions of\nthe Kitaev model to three-dimensional lattices, such as\nhyperhoneycomb and hyperoctagon lattices [44]. In the\nthree-dimensional cases, in general, the Kitaev models\nmay cause a \fnite- Tphase transition between the para-\nmagnetic and QSL phases. Indeed, such an exotic transi-\ntion was found for the hyperhoneycomb Kitaev model [9].\nThe phase transition is triggered by the con\fnement and\ndecon\fnement of emergent loops composed of excited\n\ruxes [9]. This is a topological phase transition that can-\nnot be described by a local order parameter. Although it\nis expected that dynamical quantities exhibit peculiar be-\nhavior associated with the topological phase transition,\nthe CDMFT is not able to describe such a transition.\nThus, with bearing the fact in mind that there are some\ncandidates for the three-dimensional Kitaev QSLs [45{\n49] the calculation of dynamical quantities in all Trange\nbeyond the CDMFT is an interesting challenge left for\nfuture works.\n(a) (b)FIG. 13. Schematic pictures of the di\u000berent types of clusters\nused in the benchmark of CDMFT. The color of the bonds\nare common to Fig. 1(a).\nAppendix A: Cluster size dependence\nIn the CDMFT, we replace the lattice model to the\nimpurity model with a \fnite-size cluster. The CDMFT\nbecomes exact in the limit of in\fnite size cluster. Al-\nthough the cluster size dependence was examined for the\nisotropic case with \u000b= 1:0 in Supplemental Material\nfor the previous study [18], here we present the cluster\nsize dependences of \u001fpand 1=Tp\n1for\u000b= 0:8 and 1:2 in\ncomparison with the \u000b= 1:0 case. As the onsite and\nNN-site components of 1 =Tp\n1shows almost the same T\ndependences below TH(see Fig. 9), we present only the\nonsite one.\nFigure 14 shows the cluster size dependence of \u001fpand\n1=Tp\n1obtained by the CDMFT+CTQMC calculations for\nthree di\u000berent types of clusters shown in Figs. 1(a), 13(a),\nand 13(b). In each type, we change the cluster sizes in\nthe width in the xy-chain direction while keeping that in\nthez-bond direction. This is because the width in the\nxy-chain direction is rather relevant compared to that in\nthez-bond direction in the present CDMFT, presumably\ndue to the Majorana representation based on the Jordan-\nWigner transformation along the xychains. Hereafter,\nwe de\fne the size of the cluster by the average width in\nthexy-chain direction: for instance, 4 :3 for the cluster\nin Fig. 1(a), while 4 and 5 for Figs. 13(a) and 13(b),\nrespectively.\nAs shown in Figs. 14(a)-14(j), the CDMFT+CTQMC\nresults for\u001fpshow quick convergence with respect to the\ncluster width for all the cluster types. Even close to the\narti\fcial critical temperature ~Tc, the results for the width\nlarger than 4 are almost convergent to the large width\nlimit for all types of the clusters: the remnant relative\nerrors are .5%. Note that ~Tc\u00180:014 for\u000b= 1:0,~Tc\u0018\n0:0063 for\u000b= 0:8, and ~Tc\u00180:013 for\u000b= 1:2 (for the\nrotated lattice coordinate used to calculate hSp\nj(\u001c)Sp\nj0i\nforp=x;y,~Tcbecomes slightly lower: ~Tc\u00180:0052 for\n\u000b= 0:8 and ~Tc\u00180:0094 for\u000b= 1:2).\nOn the other hand, as shown in Figs. 14(k)-14(o), the\ncluster-size dependences of 1 =T1remains up to relatively\nhigherTthan\u001fp. But the remnant relative errors are\n.10% for the cluster width larger than 4, which are su\u000e-\nciently small to observe the characteristic Tdependences\nof 1=T1as shown in Figs. 9.17\n(A)(B)(C)(a) α= 1.0\n(b) α= 0.8, p= z\n(c) α= 0.8, p= x\n(d) α= 1.2, p= z\n(e) α= 1.2, p= x\n(f) (g) (h) (i) (j) \n(k) (l) (m) (n) (o) 4.55.05.56.06.57.0\n0.270.280.290.300.31\n0.30.40.50.60.70.80.91.0\n234567891011121 / T1p\ncluster width1.51.71.92.12.32.5\n0.400.420.440.460.482.02.53.03.5\n0.500.520.540.560.58\n1.01.52.02.5\n23456789101112\ncluster width56789101112\n0.200.210.220.230.24\n0.70.80.91.01.11.21.3\n23456789101112\ncluster width(A)(B)(C)(A)(B)(C)(A)(B)(C)χpχp\n0.000.020.040.060.08\n23456789101112\ncluster width0.050.060.070.080.092025303540455055\n357911\n23456789101112\ncluster width(A)(B)(C)\nFIG. 14. Cluster-size dependences of the magnetic susceptibility \u001fpfor (a)-(e) the FM case and (f)-(j) the AFM case, and (k)-\n(o) the onsite component of the NMR relaxation rate 1 =Tp\n1: (a)(f)(k)\u000b= 1:0, (b)(c)(g)(h)(l)(m) \u000b= 0:8, and (d)(e)(i)(j)(n)(o)\n\u000b= 1:2. The data for two di\u000berent Tare plotted in each case. In (a)(f)(k), the data are common to p=zandx. Calculations\nare performed for the cluster series denoted in (A) Fig. 1(a), (B) Fig. 13(a), and (C) Fig. 13(b). Symbols in (a)-(e) are common\nfor the same parameters in (f)-(o).\nAppendix B: Accuracy of the maximum entropy\nmethod\nIn the CDMFT+CTQMC calculations, we calculate\nSp\nj;j0(!) fromhSp\nj(\u001c)Sp\nj0iby the MEM as described in\nSec. II D. In this Appendix, we examine the accuracy\nof the MEM in the limit of decoupled one-dimensional\nchains, i.e., \u000b= 1:5 (Jz= 0), where Sp\nj;j0(!) can be\nobtained directly without the MEM. We also examine\nthe accuracy by comparing Sp\nj;j0(!) at su\u000eciently low- T\nwith the analytical solution in the ground state.\nFirst, we show the comparison in the limit of decoupled\none-dimensional chains, i.e., \u000b= 1:5 (Jz= 0). In this\nlimit, the Kitaev Hamiltonian in Eq. (1) is written only\nby itinerant matter fermions fcg, in the form of Eq. (2)\nwithJz= 0. In this noninteracting problem, following\nRef. [50], we can calculate Sx\nj;j0(!) by considering the\nreal-time evolution (RTE) of hSx\nj(t)Sx\nj0i, instead of the\nimaginary-time correlation hSx\nj(\u001c)Sx\nj0i, as\nSx\nj;j0(!) =Z1\n\u00001dtei!t\u0000\u000fjtjhSx\nj(t)Sx\nj0i: (B1)\nWe call this method as the RTE in the following. In the\nRTE calculations, we consider an xychain with 600 sites\nunder the open boundary condition and take a su\u000eciently\nsmall\u000f= 0:04 in Eq. (B1).\nOn the other hand, Sz\nj;j0(!) has a nonzero value only\nfor the onsite component, which is given by 4 hSz\nj(\u001c)Sz\nji=\nhcj(\u001c)cji. Hence,Sz\nj;j(!) is obtained as\nSz\nj;j(!) =1\n2(1 +e\u0000\f!)D(!); (B2)whereD(!) is the DOS for itinerant matter fermions in\nthe one-dimensional limit:\nD(!) =1\n\u0019p\n1:52\u0000!2: (B3)\nWe call this method to estimate Sz\nj;j(!) the exact-DOS\nin the following.\nFigure 15 shows the results of Sp\nj;j0(!) obtained by the\nMEM, RTE, and exact-DOS methods for the FM case\nwith\u000b= 1:5 (Jx=Jy= 1:5 andJz= 0). We present\nboth onsite and NN-site components for Sx\nj;j0(!), while\nonly the onsite one for Sz\nj;j0(!). We \fnd that overall !\ndependence of Sp\nj;j0(!) is well reproduced by the MEM.\nIn particular, the agreement is excellent in the low !\nregion; the growth of Sx\nj;j0(!= 0) on decreasing T, which\ncontributes to 1 =T1, is well reproduced by the MEM. On\nthe other hand, the relatively sharp structures at !\u0018\n1:5 are blurred in the MEM results for both p=xand\nz, presumably because hSp\nj(\u001c)Sp\nj0iis more insensitive to\nSp\nj;j0(!) in the larger !region. Nevertheless, as shown\nin Figs. 15(e) and 15(f), the MEM results reproduce the\nbroad incoherent peak of Sx\nj;j(!)\u0000Sx\nNN(!).\nNext, we examine the accuracy of the MEM for the\ndata at su\u000eciently low Twith the analytical solution in\nthe ground state [11]. Figure 16 shows Sz(\u0000;!) obtained\nby the Majorana CDMFT+CTQMC method for the FM\ncase with\u000b= 0:8 atT= 0:003. In the ground state,\nthe energy required to \rip a single \u0011ris \u0001'0:042 at\n\u000b= 0:8. Re\recting the \rux gap, Sz(\u0000;!) at lowThas\na\u000e-function like peak at \u0001 '0:042 [11]. As shown in\nFig. 16, our CDMFT+CTQMC result shows a peak at\nthis energy, which is considered to precisely reproduce18\nSx\njj(ω)\n-0.5 Sx\nNN(ω) Sz\nj,j(ω)\nω ω(a) (b)\n(c) (d)\n(e) (f)\n(g) (h)0.00.51.0\n0.00.51.0\n0.00.10.20.30.4\n0.00.20.40.60.8\n-1 0 1 -1 0 1 2MEM\nRTEMEM\nRTE\nMEM\nRTEMEM\nRTE\nMEM\nRTEMEM\nRTE\nMEM\nexactMEM\nexactS(ω) -Sx\nNN(ω)x\nj,j\nFIG. 15. Comparison between the MEM, RTE, and exact-\nDOS results for (a)(b) Sx\nj;j(!), (c)(d)Sx\nNN(!), (e)(f)Sx\nj;j(!)\u0000\nSx\nNN(!), and (g)(h) Sz\nj;j(!) at (a)(c)(e)(g) T= 0:375 and\n(b)(d)(f)(h) T= 0:0375.\nthe low-energy structure of the dynamical spin structure\nfactor.\nFrom these observations, we consider that the MEM\nresults for S(q;!) and 1=T1in Sec. III B and III C are\naccurate enough to discuss the Tand!dependences.\nAppendix C: Spin correlations as functions of Tand\n!\nIn this Appendix, we present the spin correlations\nas functions of Tand!, which are obtained by the\nMEM. Figure 17 shows the results for onsite and NN-site\ncomponents for \u000b= 1:0, 0:8, and 1:2. The data are used\nto obtain the dynamical quantities in Sec. III B and III C.\n0123456\n0.000.020.040.060.080.10Sz(Γ, ω)\nωFIG. 16. Sz(\u0000;!) obtained by the Majorana\nCDMFT+CTQMC method for the FM case for \u000b= 0:8 at\nT= 0:003. Vertical line at !\u00180:042 represents the value of\n\rux gap of the ground state calculated exactly.\nAppendix D: Tdependence of the Korringa ratio\nFigures 18 and 19 display the Tdependences of the\nKorringa ratio de\fned as\nKp=1\nTp\n1T(\u001fp)2; (D1)\nwhich is computed by using the NMR relaxation rate\n1=Tp\n1and the magnetic susceptibility \u001fpobtained in\nSec. III C and III D. Interestingly, as shown in Fig. 18(a),\nKpfor the isotropic FM case is almost constant close to\n1 forTL.T.TH, which is apparently consistent with\nthe behavior expected for free electron systems. This is\nalso the case for the xcomponent for the FM case with\n\u000b= 1:2, as shown in Fig. 18(c). However, the suggestive\nbehavior is presumably super\fcial, as the results for the\nAFM cases as well as for \u000b= 0:8 behave di\u000berently with\nsubstantial Tdependence.\nACKNOWLEDGMENTS\nThe authors thank M. Imada, Y. Kamiya, K. Oh-\ngushi, S. Takagi, M. Udagawa, and Y. Yamaji for\nfruitful discussions. Y. M. thanks A. Banerjee, C.\nD. Batista, K.-Y. Choi, S. Ji, S. Naglar, and J.-H.\nPark for constructive suggestions. This research was\nsupported by Grants-in-Aid for Scienti\fc Research un-\nder Grants No. JP15K13533, No. JP16K17747, and\nNo. JP16H02206. Parts of the numerical calculations\nwere performed in the supercomputing systems in ISSP,\nthe University of Tokyo.\n[1] D. C. Tsui, H. L. Stormer, and A. C. Gossard, Two-\nDimensional Magnetotransport in the Extreme QuantumLimit, Phys. Rev. Lett. 48, 1559 (1982).19\n(a) (b)\n(c) (d)\n(e) (f)\n(g) (h)\n(i) (j)\nFIG. 17. Spin correlations as functions of Tand!for\nthe FM case: (a)(c)(e)(g)(i) onsite components Sj;j(!) and\n(b)(d)(f)(h)(j) NN-site components SNN(!). (a) and (b) are\nforSx\nj;j(!) =Sz\nj;j(!) andSx\nNN(!) =Sz\nNN(!), respectively, at\n\u000b= 1:0. (c), (d), (e), and (f) [(g), (h), (i), and (j)] are for\nSx\nj;j(!),Sz\nNN(!),Sz\nj;j(!), andSz\nNN(!), respectively, at \u000b= 0:8\n(1:2). The white and gray dotted lines indicate THandTL,\nrespectively, for each \u000b.\n[2] H. L. Stormer, D. C. Tsui, and A. C. Gossard, The frac-\ntional quantum Hall e\u000bect, Phys. Mod. Phys. 71, S298\n(1999).\n[3] P. W. Anderson, Resonating valence bonds: A new kind\nof insulator?, Mater. Res. Bull. 8, 153 (1973).\n[4] S. A. Kivelson, D. S. Rokhsar, and J. P. Sethna, Topology\nof the resonating valence-bond state: Solitons and high-\nTcsuperconductivity, Phys. Rev. B 35, 8865 (1987).\n[5] T. Senthil and M. P. A. Fisher, Z2gauge theory of\nelectron fractionalization in strongly correlated systems,\nPhys. Rev. B 62, 7850 (2000).\n[6] M. Hermele, M. P. A. Fisher, and L. Balents, Pyrochlore\nphotons: The U(1) spin liquid in a S=1\n2three-\ndimensional frustrated magnet, Phys. Rev. B, 69, 064404\nT(a)\n(b)\n(c)zx\nzx\n10-1100101\n10-210-110010-1100101\n10-410-310-210-1100101102onsite\nNN-site\nonsite\nNN-site\nonsite\nNN-siteFIG. 18. Tdependences of the Korringa ratio Kp=\n1=(Tp\n1T(\u001fp)2) for the FM case at (a) \u000b= 1:0, (b)\u000b= 0:8, and\n(c)\u000b= 1:2 (p=z;x). Note thatKz=Kxfor\u000b= 1:0 and\nKx=Kyfor all the cases from the symmetry. The vertical\ndotted lines indicates TLandTHfor each\u000b.\n(2004).\n[7] C. Castelnovo, R. Moessner, and S. L. Sondhi, Magnetic\nmonopoles in spin ice, Nature 451, 42 (2008).\n[8] A. Kitaev, Anyons in an exactly solved model and be-\nyond, Ann. Phys. 321, 2 (2006).\n[9] J. Nasu, M. Udagawa, and Y. Motome, Vaporization\nof Kitaev Spin Liquids, Phys. Rev. Lett. 113, 197205\n(2014).\n[10] J. Nasu, M. Udagawa, and Y. Motome, Thermal frac-\ntionalization of quantum spins in a Kitaev model:\nTemperature-linear speci\fc heat and coherent transport\nof Majorana fermions, Phys. Rev. B 92, 115122 (2015).\n[11] J. Knolle, D.L. Kovrizhin, J.T. Chalker, and R. Moess-\nner, Dynamics of a Two-Dimensional Quantum Spin Liq-\nuid: Signatures of Emergent Majorana Fermions and\nFluxes, Phys. Rev. Lett. 112, 207203 (2014).\n[12] A. Banerjee, C. A. Bridges, J-Q. Yan, A. A. Aczel, L. Li,\nM. B. Stone, G. E. Granroth, M. D. Lumsden, Y. Yiu, J.20\nT(a)\n(b)\n(c)zx\nzx\n10-1100101102103\n10-210-110010-110010110210310-1100101102103104\nonsite\nNN-site\nonsite\nNN-site\nonsite\nNN-site\nFIG. 19. Tdependences of the Korringa ratio Kp=\n1=(Tp\n1T(\u001fp)2) for the AFM case at (a) \u000b= 1:0, (b)\u000b= 0:8,\nand (c)\u000b= 1:2 (p=z;x). The notations are common to\nthose in Fig. 18.\nKnolle, D. L. Kovrizhin, S. Bhattacharjee, R. Moessner,\nD. A. Tennant, D. G. Mandrus, S. E. Nagler, Proximate\nKitaev quantum spin liquid behaviour in a honeycomb\nmagnet, Nature Mater. 15733 (2016).\n[13] S.-H. Do, S.-Y. Park, J. Yoshitake, J. Nasu, Y. Motome,\nY. S. Kwon, D. T. Adroja, D. J. Voneshen, K. Kim,\nT.-H. Jang, J.-H. Park, K.-Y. Choi, S. Ji, Incarnation\nof Majorana Fermions in Kitaev Quantum Spin Lattice,\npreprint (arXiv:1703.01081).\n[14] J. Knolle, G.-W. Chern, D. L. Kovrizhin, R. Moessner,\nand N. B. Perkins, Raman Scattering Signatures of Ki-\ntaev Spin Liquids in A2IrO3Iridates with A=Na or Li,\nPhys. Rev. Lett. 113, 187201 (2014).\n[15] L. J. Sandilands, Y. Tian, K. W. Plumb, Y.-J. Kim, and\nK. S. Burch, Scattering Continuum and Possible Frac-\ntionalized Excitations in \u000b-RuCl 3, Phys. Rev. Lett. 114,\n147201 (2015).\n[16] J. Nasu, J. Knolle, D. L. Kovrizhin, Y. Motome, and R.\nMoessner, Fermionic response from fractionalization inan insulatin two-dimensional magnet, Nature Phys. 12,\n912 (2016).\n[17] A. Glamazda, P. Lemmens, S.-H. Do, Y. S. Choi, and K.-\nY. Choi, Raman spectroscopic signature of fractionalized\nexcitations in the harmonic-honeycomb iridates \fand\r-\nLi2IrO3, Nat. Commun. 7, 12286 (2016).\n[18] J. Yoshitake, J. Nasu, and Y. Motome, Fractional Spin\nFluctuation as a Precursor of Quantum Spin Liquids:\nMajorana Dynamical Mean-Field Study for the Kitaev\nModel, Phys. Rev. Lett. 117, 157203 (2016).\n[19] R. D. Johnson, S. C. Williams, A. A. Haghighirad, J.\nSingleton, V. Zapf, P. Manuel, I. I. Mazin, Y. Li, H. O.\nJeschke, R. Valent\u0013 \u0010, and R. Coldea, Monoclinic crystal\nstructure of \u000b\u0000RuCl 3and the zigzag antiferromagnetic\nground state, Phys. Rev. B, 92, 235119 (2015).\n[20] Y. Yamaji, Y. Nomura, M. Kurita, R. Arita, and M.\nImada, First-Principles Study of the Honeycomb-Lattice\nIridates Na 2IrO3in the Presence of Strong Spin-Orbit\nInteraction and Electron Correlations, Phys. Rev. Lett.\n113, 107201 (2014).\n[21] S. K. Choi, R. Coldea, A. N. Kolmogorov, T. Lancaster,\nI. I. Mazin, S. J. Blundell, P. G. Radaelli, Yogesh Singh,\nP. Gegenwart, K. R. Choi, S.-W. Cheong, P. J. Baker,\nC. Stock, and J. Taylor, Spin Waves and Revised Crys-\ntal Structure of Honeycomb Iridate Na 2IrO3, Phys. Rev.\nLett. 108, 127204 (2012).\n[22] S. M. Winter, Y. Li, H. O. Jeschke, and R. Valenti, Chal-\nlenges in design of Kitaev materials: Magnetic interac-\ntions from competing energy scales, Phys. Rev. B 93,\n214431 (2016).\n[23] G. Baskaran, Saptarshi Mandal, and R. Shankar, Exact\nResults for Spin Dynamics and Fractionalization in the\nKitaev Model, Phys. Rev. Lett. 98, 247201 (2007).\n[24] S. Mandal, R. Shankar, and G. Baskaran, RVB gauge the-\nory and the topological degeneracy in the honeycomb Ki-\ntaev model, J. Phys. A Math. Theor. 45, 335304 (2012).\n[25] H.-D. Chen and J. Hu, Exact mapping between classical\nand topological orders in two-dimensional spin systems,\nPhys. Rev. B 76, 193101 (2007).\n[26] X.-Y. Feng, G.-M. Zhang, and T. Xiang, Topological\nCharacterization of Quantum Phase Transitions in a\nSpin-1=2 Model, Phys. Rev. Lett. 98, 087204 (2007).\n[27] H.-D. Chen, and Z. Nussinov, Exact results of the Ki-\ntaev model on a hexagonal lattice: spin states, string\nand brane correlators, and anyonic excitations, J. Phys.\nA Math. Theor. 41, 075001 (2008).\n[28] G. Kotliar, S. Y. Savrasov, G. P\u0013 alsson, and G. Biroli,\nCellular Dynamical Mean Field Approach to Strongly\nCorrelated Systems, Phys. Rev. Lett. 87, 186401 (2001).\n[29] N. Furukawa, Transport Properties of the Kondo Lattice\nModel in the Limit S=1andD=1, J. Phys. Soc.\nJpn.63, 3214 (1994).\n[30] J. Nilsson and M. Bazzanella, Majorana fermion descrip-\ntion of the Kondo lattice: Variational and path integral\napproach, Phys. Rev. B 88, 045112 (2013).\n[31] W. Metzner and D. Vollhardt, Correlated Lattice\nFermions in d=1Dimensions, Phys. Rev. Lett. 62,\n324 (1989).\n[32] A. Georges, G. Kotliar, W. Krauth, and M. J. Rozen-\nberg, Dynamical mean-\feld theory of strongly correlated\nfermion systems and the limit of in\fnite dimensions, Rev.\nMod. Phys. 68, 13 (1996).\n[33] M. Udagawa, H. Ishizuka, and Y. Motome, Non-Kondo\nMechanism for Resistivity Minimum in Spin Ice Conduc-21\ntion Systems, Phys. Rev. Lett. 108, 066406 (2012).\n[34] P. Werner, A. Comanac, L. de' Medici, M. Troyer, and A.\nJ. Mills, Continuous-Time Solver for Quantum Impurity\nModels, Phys. Rev. Lett. 97, 076405 (2006).\n[35] Z.-X. Li, Y.-F. Jiang, and H. Yao, Solving the fermion\nsign problem in quantum Monte Carlo simulations by\nMajorana representation, Phys. Rev. B 91, 241117(R)\n(2015).\n[36] A. N. Rubtsov, V. V. Savkin, and A. I. Lichten-\nstein, Continuous-time quantum Monte Carlo method for\nfermions, Phys. Rev. B 72, 035122 (2005).\n[37] L. Boehnke, H. Hafermann, M. Ferrero, F. Lechermann,\nO. Parcollet, Orthogonal polynomial representation of\nimaginary-time Green's functions, Phys. Rev. B 84,\n075145 (2011).\n[38] R. Levy, J. P. F. LeBlanc, and E. Gull, Implementation of\nthe maximum entropy method for analytic continuation,\nComp. Phys. Commun. 215, 149 (2017).\n[39] M. Jarrell and J. E. Gubernatis, Bayesian inference and\nthe analytic continuation of imaginary-time quantum\nMonte Carlo data, Phys. Rep. 269, 133 (1996).\n[40]Sp\nNN(!) for the FM case ( Sp\nNN(!)FM) and the AFM\ncase (Sp\nNN(!)AFM) satisfy the relation Sp\nNN(!)FM =\n\u0000Sp\nNN(!)AFM. Nevertheless, Sp\nNN(!)FMandSp\nNN(!)AFM\nobtained by the present numerical procedure violates the\nrelation because of the precision of the MEM. In the\ndata presented in Sec. III, we calculate Sp\nNN(!)FMby\nthe procedure, and Sp\nj;j0(!)AFM are set asSp\nj;j0(!)AFM =\n\u0000Sp\nj;j0(!)FM.\n[41] T. Moriya, Nuclear Magnetic Relaxation in Antiferro-\nmagnetics, II, Prog. Theor. Phys., 16, 641 (1956).\n[42]\u001fxfor the AFM case at \u000b= 0:8 also shows similar behav-\nior, but it is qualitatively explained by the Curie-Weiss\nlaw in Eq. (45) for this case.\n[43] Thermal averages h\u0011r0iobtained by the CTQMC calcu-\nlations deviate from 1 even in the uniform case, althoughwe \fxP(\u0011r0= 1) = 1 in the CDMFT. This is because\nthe statistical weight of \u0011r0=\u00001 is nonzero in the two-\nsite impurity problem in the CTQMC calculations. Such\na discrepancy ofh\u0011r0ibetween the CDMFT and CTQMC\nsolutions never occurs when we do not \fx P(\u0011r0) by hand.\n[44] M. Hermanns and S. Trebst, Quantum spin liquid with a\nMajorana Fermi surface on the three-dimensional hyper-\noctagon lattice, Phys. Rev. B 89, 235102, (2014).\n[45] Y. Okamoto, M. Nohara, H. Aruga-Katori, and H. Tak-\nagi, Spin-Liquid State in the S= 1=2 Hyperkagome\nAntiferromagnet Na 4Ir3O8, Phys. Rev. Lett. 99, 137207\n(2007).\n[46] H. Kuriyama, J. Matsuno, S. Niitaka, M. Uchida, D.\nHashizume, A. Nakao, K. Sugimoto, H. Ohsumi, M.\nTakata, and H. Takagi, Epitaxially stabilized iridium\nspinel oxide without cations in the tetrahedral site, Appl.\nPhys. Lett. 96182103 (2010).\n[47] K. A. Modic, T. E. Smidt, I. Kimchi, N. P. Breznay, A.\nBi\u000en, S. Choi, R. D. Johnson, R. Coldea, P. Watkins-\nCurry, G. T. McCandless, J. Y. Chan, F. Gandara, Z.\nIslam, A. Vishwanath, A. Shekhter, R. D. McDonald, and\nJ. G. Analytis, Realization of a three-dimensional spin-\nanisotropic harmonic honeycomb iridate, Nat. Comm. 5,\n4203 (2014).\n[48] T. Takayama, A. Kato, R. Dinnebier, J. Nuss, H. Kono,\nL. S. I. Veiga, G. Fabbris, D. Haskel, and H. Takagi, Hy-\nperhoneycomb Iridate \f-Li2IrO3as a Platform for Kitaev\nMagnetism, Phys. Rev. Lett. 114, 077202 (2015).\n[49] I. Kimchi and A. Vishwanath, Kitaev-Heisenberg models\nfor iridates on the triangular, hyperkagome, kagome, fcc,\nand pyrochlore lattices, Phys. Rev. B 89, 014414 (2014).\n[50] O. Derzhko, T. Krokhmalskii, and J. Stolze, Dynamics\nof the spin-1\n2isotropicXYchain in a transverse \feld, J.\nPhys. A: Math. Gen., 33, 3063 (2000)." }, { "title": "1609.02319v1.Pumping_dynamics_of_nuclear_spins_in_GaAs_quantum_wells.pdf", "content": "arXiv:1609.02319v1 [cond-mat.mes-hall] 8 Sep 2016Pumping dynamics of nuclear spins in GaAs quantum wells\nRaphael W. Mocek,∗Danil O. Tolmachev, Giovanni Cascio, and Dieter Suter†\nExperimental Physics III, TU Dortmund University, Otto-Ha hn-Str. 4a, 44227 Dortmund,Germany\n(ΩDated: August 3, 2018)\nIrradiating a semiconductor with circularly polarized lig ht creates spin-polarized charge carriers.\nIf the material contains atoms with non-zero nuclear spin, t hey interact with the electron spins via\nthe hyperfinecoupling. Here, we consider GaAs/AlGaAs quant um wells, where the conduction-band\nelectron spins interact with three different types of nuclea r spins. The hyperfine interaction drives\na transfer of spin polarization to the nuclear spins, which t herefore acquire a polarization that is\ncomparable to that of the electron spins. In this paper, we an alyze the dynamics of the optical\npumping process in the presence of an external magnetic field while irradiating a single quantum\nwell with a circularly polarized laser. We measure the time d ependence of the photoluminescence\npolarization to monitor the buildup of the nuclear spin pola rization and thus the average hyperfine\ninteraction acting on the electron spins. We present a simpl e model that adequately describes the\ndynamics of this process and is in good agreement with the exp erimental data.\nPACS numbers: 78.67.De, 78.55.Cr, 78.66.Fd, 76.60.-k\nKeywords: Optical Orientation, Semiconductors, Optical P umping, Spin Dynamics\nI. INTRODUCTION\nOptical pumping creates electrons and holes in semi-\nconductor samples with spin polarizations far from equi-\nlibrium, as shown by the level scheme and transition dia-\ngramoffig. 1. Dependingontheconduction-andvalence-\nband states involved in the optical transition, the polar-\nization of the electron spins can reach almost 100%[ 1].\nFIG. 1. Selection rules for the optical transitions between\nthe valence-band (VB) and the conduction-band (CB) of a\nsemiconductor quantum well. The confinement lifts the de-\ngeneracy of the hole states. (modified after[ 2])\nThe spin-polarizationofthe conduction-bandelectrons\ncan be monitored through the photoluminescence (PL)\npolarization[ 3–5]\nDOP=I(σ+)−I(σ−)\nI(σ+)+I(σ−), (1)\nwhereI(σ±) is the intensity of right- or left-circularly\npolarized PL.\n∗raphael.mocek@tu-dortmund.de\n†dieter.suter@tu-dortmund.deInmanymaterials,andinparticularintheGaAsquan-\ntum wells that we consider in this work, the electrons are\ncoupled to different nuclear spins by the hyperfine inter-\naction. Accordingly, the electron spin orientation can be\ntransferred to the nuclear spins[ 6–9] in a process known\nas dynamic nuclear polarization (DNP)[ 2,10,11]. Con-\nversely, the ensemble of polarized nuclear spins affect the\nevolution of the electron spins. The overall effect can be\nsummarizedbyaneffectivenuclearmagneticfield[ 10,12–\n17]. This is used, e.g., in optically detected nuclear mag-\nnetic resonance[ 6,13,18–22].\nThe goal of this study is a detailed understanding of\nthe buildup of the nuclear spin polarization during opti-\ncal pumping. For this purpose, we rely mostly on mea-\nsurements ofthe time dependence of the PL polarization,\nfrom which we determine the buildup of the nuclearmag-\nnetic field. The paper is organized as follows. In Sec. II,\nwe describe a simple model of the spin dynamics during\noptical pumping. In Sec. IIIwe present the experimental\nsetup and the sample under investigation. Section IVA\ncontainstheexperimentalresultsforthetimedependence\nofthe optical pumping processandcomparesthem to the\ntheoretical prediction. Section IVBsummarizes the ef-\nfect of the control parameters laser intensity and optical\ndetuning on the optical pumping dynamics. The paper\nends with a short discussion and conclusions.\nII. THEORY\nA. Electron spin polarization\nThe optical pumping process, as well as the optical\ndetection couple the photon angular momentum directly\nto the spin ofthe chargecarriers. We therefore start with\nthe equation of motion for the spin density operator ρ:2\n∂ρ\n∂t=−i\n/planckover2pi1[H,ρ]−ΓRρ−ΓS/parenleftbigg\nρ−Tr{ρ}\n2\n/BD/parenrightbigg\n+˜P(2)\nH=/planckover2pi1γe/vectorB·/vectorS\nWe use the spin operators defined as Si=1\n2σiwithi∈\n[x,y,z],γeis the gyromagnetic ratio of the conduction\nelectrons and /BDis the two-dimensional unity matrix. /vectorB\nis the magnetic field, Γ Rdescribes the recombination of\nthe electrons from the conduction- to the valence-band\nand ΓSthe spin relaxation rate. The matrix ˜Pdescribes\nthe buildup of electron spin density by the absorption of\ncircularly polarized photons. In the coordinate system\ndefined in fig. 2, this matrix is\n˜P=P/parenleftbigg\ncos2ΘL\n21\n2sinΘL\n1\n2sinΘLsin2ΘL\n2/parenrightbigg\n. (3)\nPis the rate at which the optical pumping process gener-\nates electron spin density in the conduction band. Θ Lis\nthe angle between the incident laser beam and the z-axis,\nas defined in fig. 2.\nFIG. 2. Schematic overview of the chosen coordinate system\nand the relevant angles. The laser light hits the sample per-\npendicularly.\nIf the nuclear spins are polarized, they collectively\nmodify the effective magnetic field acting on the electron\nspin. We therefore write the total field /vectorBas the sum\nof the external field /vectorBextand the nuclear field /vectorBnuc. In\nour coordinate system, both are oriented along the z-axis\nand we therefore write Bext+Bnucfor thez−component\nof the effective magnetic field. Under stationary condi-\ntions, the expectation values of the three components of\nthe electron spin are\n/an}bracketle{tSx/an}bracketri}ht=Tr{Sxρ}=P\n2γe∆BsinΘL\n(Bext+Bnuc)2+∆B2(4)\n/an}bracketle{tSy/an}bracketri}ht=Tr{Syρ}=P\n2γe(Bext+Bnuc)sinΘ L\n(Bext+Bnuc)2+∆B2\n/an}bracketle{tSz/an}bracketri}ht=Tr{Szρ}=P\n2γecosΘL\n∆B.\nHere the parameter ∆ Bis defined as ∆ B=/planckover2pi1(ΓR+ΓS)\n|g∗|µB\nwithg∗asthe g-factorofthe conduction electrons[ 23–25]andTr{ρ}=P\nΓR. In our case, the direction of detection\n/vector eis close to the x-axis\n/vector e=/parenleftbigsinΘD,0,cosΘD/parenrightbig\n, (5)\nwhere the angle θDis defined in fig. 2.\nExperimentally, we measure the degree of photon po-\nlarization (see eq.( 1)) in the direction /vector eby dividing the\ndifferencebetweenthetwointensities I(σ±)bytheirsum.\nThe result is\nSD=1\nTr{ρ}(sinΘD/an}bracketle{tSx/an}bracketri}ht+cosΘ D/an}bracketle{tSz/an}bracketri}ht)\n=S0/parenleftBigg\ncosΘLcosΘD+∆B2sinΘLsinΘD\n∆B2+(Bext+Bnuc)2/parenrightBigg\n.(6)\nHereS0=1\n2ΓR\nΓR+ΓSdescribes the equilibrium spin polar-\nization in the absence of a magnetic field. This signal,\nmeasured as a function of the external magnetic field,\nis known as a (shifted) Hanle curve. It represents a\nLorentzian, with a maximum at Bext=−Bnucand a\nwidth ∆B.\nB. Nuclear spin polarization\nIn this study, we concentrate on the evolution of the\nnuclear spin polarization. The equation of motion for the\nnuclear spin populations can be written as\nd\ndt/parenleftbigg\np↑\np↓/parenrightbigg\n=/parenleftbigg−κs↓−1\n2T1κs↑+1\n2T1\nκs↓+1\n2T1−κs↑−1\n2T1/parenrightbigg/parenleftbigg\np↑\np↓/parenrightbigg\n,(7)\nwhereκis the transfer rate at which electronic and nu-\nclear spins undergo mutual flip-flop transitions and s↑↓\nare the densities of the electron spins generated by the\noptical pumping process. According to eq.( 4) they are\ns↑↓=P\n2ΓR±/an}bracketle{tSz/an}bracketri}ht.\nFor quantum wells with dimensions of ≈20 nm, the\nrecombination rate of the electrons is Γ R≈109s−1[26].\nFor time-independent parameters, eq.( 7) can be solved\nanalytically. If the system is initially in thermal equilib-\nrium,p↑↓(0) = 0.5, the solution is\np↑↓(t) =1±∆p(t)\n2, (8)\nwhere the population difference is\n∆p(t) = ∆p∞/parenleftbigg\n1−e−/parenleftBig\n1\nT1+κP\nΓR/parenrightBig\nt/parenrightbigg\n(9)\nand it’s equilibrium value\n∆p∞=/an}bracketle{tSz/an}bracketri}ht2κT1ΓR\nΓR+κT1P. (10)3\nThe nuclear spin polarization can be measured through\nits effect on the electron spin, via the effective nuclear\nfield\nBnuc(t) =Bmax∆p(t)\n=Bmax∆p∞/parenleftbigg\n1−e−/parenleftBig\n1\nT1+κP\nΓR/parenrightBig\nt/parenrightbigg\n.(11)\nAccording to eq.( 6),Bnucis given by the maximum of\nthe Hanle curve. To measure the time dependence of the\npopulations of the nuclear spin, we therefore measure the\nHanle curves for different pumping times.\nIII. EXPERIMENTAL\nThe sample used for this investigation was grown by\nmolecular beam epitaxy on a Te-doped GaAs substrate.\nIt consists of 13 undoped GaAs/Al0.35Ga0.65As quantum\nwells with thicknesses dranging from 2.8 to 39 .3nm[27].\nFIG. 3. Experimental setup of the optical pumping exper-\niment. L1, L2 and L3: lenses, LP: linear polarizer, PEM:\nphoto elastic modulator, APD: avalanche photo diode, BS:\nbeam splitter, λ/2,λ/4 : retardation plates.\nFigure3shows a schematic representation of the experi-\nmental setup. For the optical excitation we use a semi-\nconductor laser (Toptica DLC DL PRO), which covers\nthe wavelength range of λexc= 799−812nm. The mag-\nneticfieldiscreatedbyaresistiveelectromagnet(Bruker)\nwith a range of Bext= 0T−1.4T. The sample is\nmounted on the cold finger of a home-built flow-cryostat\nand kept at temperatures of T≈4.7±0.3K. The laser\nbeam analysis includes a spectrometer (APE waveScan\nUSB) for monitoring the laser wavelength and a pho-\ntodiode to monitor the laser power. The PL is passed\nthrough a monochromator (Spex 1704) and a photo-\nelastic modulator (Hinds Instruments PEM 90) and de-\ntected with an avalanche photodiode (APD, Hamamatsu\n5640). Two lock-in amplifiers (Stanford Research SR830\nDSP) are then used to measure the total PL powerIΣ=I(σ+) +I(σ−) and the difference between right-\nand left-circularly polarized light I∆=I(σ+)−I(σ−).\nThe optical pumping process took place in constant\nexternal magnetic fields of Bext= 0.3T or 1T. We\nmonitored the buildup of the nuclear spin polarization\nby measuring Hanle curves[ 2,28] as a function of the\npumping time. For the Hanle curves, the magnetic field\nwas scanned from Bexteither upward or downward, de-\npending on the displacement of the Hanle curve. The\ntime for measuring a Hanle curve was about 10s, short\ncompared to the duration of the optical pumping. Dur-\ningtheHanlemeasurements,the laserpowerwasreduced\ntoPL≈2mW, to minimize the optical pumping effects.\nThe angles defined in fig. 2were Θ D= 81◦and Θ L= 78◦\nfor all experiments.\nIV. RESULTS\nA. Nuclear field buildup\nThe main goal of these experiments was a quantita-\ntive understanding of the process that generates the nu-\nclear spin polarization. For this purpose, we performed\na set of measurements that consisted of a pumping pe-\nriodTpumpduring which the sample was irradiated with\ncircularly polarized light in a constant magnetic field.\nImmediately after this pumping period, we performed\na rapid scan of the magnetic field to measure a Hanle\ncurve. According to eq.( 4), the maxima of these curves\ncorrespond to the effective nuclear field |Bnuc|and can\ntherefore be used as a probe of the nuclear spin polar-\nization. Measured Hanle curves after two different times\nTpumpare shown in fig. 4(a). The experimental param-\neters for these measurements are Bext= 1T and laser\npowerPL= 49mW. The monochromator was set to\nthe maximum of the PL line of the d= 19.7nm quan-\ntum well and the optical detuning of the laser beam was\n∆λ=λdet−λexc= 811.6nm−811.3nm = 0 .3nm.\nFigure4(b) shows the buildup of the effective nuclear\nfield|Bnuc(t)|. It compares the experimental data with\nthe theoretical curve calculated from eq.( 11). For the\nparameters,weusedaspin-latticerelaxationtimeof T1=\n596±160s, which we measured independently, and is\ncomparable to literature values for similar systems [ 29].\nFrom the fitted curve and our experimental parameters,\nwe calculated P= 7.7·1013#e−\ns·µm3. Table Ishows the\nother relevant parameters determined from these curves.\nBmax[T] ∆ p∞[%]κ/bracketleftBig\nµm3\ns·#e−/bracketrightBig\n29 3 .4 6.8·10−8\nTABLE I. Fit parameters for the data shown in fig. 4(b).\nThe fit-result for the maximum field Bmaxis signifi-\ncantly larger than some values from the literature [ 2,12].\nThebuildupofthenuclearspinpolarizationcanalsobe4\n(b)\nFIG. 4. Time dependence of the effective nuclear field |Bnuc|:\n(a) Hanle curves measured after optical pumping periods of\nTpump= 120s and 900s. (b) Evolution of the nuclear field\nduring the optical pumping. The solid red line is the fit resul t\nusing eq.( 11) and the filled circles represent the experimental\nvalues.\nmonitored through the time dependence of the PL polar-\nization during the optical pumping process, as shown in\nfig.5. The theoretical curve was calculated from eqs.( 6)\nand (11) with the parameters of tab. I.\nB. Dependence on the laser intensity\nThe parameter Pintroduced in eq.( 3) describes the\nrateat which electron spin density is createdin the mate-\nrial. Over some range, we therefore expect that Pshould\nbe proportional to the laser power PL. Since only ab-\nsorbed photons generate electron spins, the rate should\nalso depend on the absorption probability of the photons\nand reach a maximum at the optical resonance. The in-\nfluence of optical detuning is discussed in Sec. IVC.\nWe examined the dependence of the pumping dy-\nnamics on the laser power by performing a series of\nmeasurements with increasing laser intensity. After\nFIG. 5. Time dependence of the optical pumping process:\nmeasured PL polarization under optical pumping conditions\nwithBext= 1T,PL= 49mW and ∆ λ= 0.3nm. The solid\nred line was calculated with the parameters of table Iusing\neq.(6) and (11).\npumping times Tpump= [30s,60s,180s,300s,600s],\nwe measured Hanle curves to monitor the evolution\nof the effective nuclear magnetic field |Bnuc|. We\nrepeated this procedure with laser powers of PL=\n[20mW,25mW,30mW,35mW,40mW]. Further ex-\nperimental parameters were Bext= 0.3T and ∆ λ=\n0.5nm. The detection wavelength was λdet= 811.6nm,\nwhich corresponds to the maximum of the PL line of the\nd= 19.7nm quantum well.\nFIG. 6. Evolution of the effective nuclear magnetic field\n|Bnuc|for different laser powers PL. The circles mark the\nexperimental data while the solid lines are the result of the\nfit using eq.( 11) andBmax= 5.3T,κ= 6.3·10−8µm3\ns·#e−as\nfixed parameters.\nFigure6showsthebuildupoftheeffectivenuclearmag-5\nnetic field |Bnuc(t)|for different laser intensities. The\nexperimental data, which are shown as circles, are the\nmaxima of the Hanle curves taken after each optical\npumping period Tpump. We used Bmax= 29T and\nκ= 6.8·10−8µm3\ns·#e−as fixed parameters obtained by\nthe measurement presented in Sec. IVAand eq.( 11) to\nfit the experimental data shown in fig. 6. The resulting\n|Bnuc(t)|are shown as solid curves.\nFIG. 7. Rate of change in electron spin density Ptogether\nwith the PL light power as a function of the laser power PL.\nThe solid line is the result of the linear fit.\nFrom the observed buildup-rate of |Bnuc|, we calcu-\nlated the rate Pat which electron-spin density is gen-\nerated by the pumping process. Figure 7shows the re-\nsulting rates Pas a function of the laser power PL. The\nrate at which electrons are generated in the conduction-\nband should also be reflected in the PL power, which we\nmeasured independently. As shown in fig. 7, both quan-\ntities are roughly proportional to the laser power, with\nproportionality factors mP= (1.7±0.5)·1012#e−\ns·µm3·mW\nandmPLpower= 0.4±0.1nW\nmW, respectively.\nC. Optical detuning\nThe rate Pat which electron-spins are generated de-\npends also on the frequency of the laser with respect to\nthe resonance frequency of the quantum well. We mea-\nsured the time dependence of the PL polarization un-\nder optical pumping conditions for different optical de-\ntunings ∆ λ= [0.7...2.1]nm relative to the peak wave-\nlengthλdet= 811.6nm of the d= 19.7nm quantum\nwell. The experimental parameters were Bext= 0.3T\nandPL= 47mW.\nFigure8shows some of the measured curves. The ex-\nperimental data are compared to the theoretical expec-\ntations calculated from eq.( 6) and (11), using the exper-\nimental parameters of our system and the fit parameters\nFIG. 8. Time dependence of the optical pumping process:\nmeasured PL polarization under optical pumping conditions\nwithBext= 0.3T,PL= 47mW and Tpump= 600s for differ-\nent optical detunings ∆ λ. The solid lines are the fit results\nbased on eq.( 11),(6).\ngiven in tab. I. The electron-spin density generation rate\nPwas adjusted for the curves to fit the experimental\ndata.\nFIG. 9. Detuningdependence of the rate of change in electron\nspin density P: the circles mark the fit-parameters Pand\nthe solid red line is the result of the Lorentzian fit function\neq.(12). Detuning dependence of the PL light power: the\ndiamonds mark the measured PL light power and the solid\nblack line is given by eq.( 13).\nFigure9shows the resulting fit-parameters Pas red cir-\ncles for the complete set of measurements as a function\nof the laser detuning ∆ λ.We compare the experimental6\ndata points to a Lorentzian,\nP(∆λ) =a1/parenleftbig∆λ\nHWHM/parenrightbig2+1. (12)\nFitting the data marked as circles in fig. 9to eq.(12) re-\nsults ina1= (5.5±0.9)·1015#e−\ns·µm3·nmand a half width\nat half maximum HWHM = (0.5±0.1) nm.\nThe number of photons absorbed by the QW should\nalso be reflected in the rate of emitted PL. We there-\nfore also measured the average PL power as a function\nof the detuning ∆ λ. The results are shown in fig. 9as\nblack diamonds. We compare them to the theoretically\nexpected behavior of a Lorentzian line similar to eq.( 12),\nbut with an additional offset PLoffreflecting the effect\nof non-resonant excitation:\nPLpower(∆λ) =a2/parenleftbig∆λ\n0.5nm/parenrightbig2+1+PLoff.(13)\nThe fit result is shown in fig. 9as a solid black line using\nthe fit-parameters PLpower(∆λ= 0)≈40.8nW,a2=\n31.7±2.1nW and PLoff= 9.2±0.8nW. The value of\nPLoffwas also measured independently by exciting the\nsystem with a blue laser with λexc= 406nm and a laser\npower of PL= 10mW. We obtained a PL power that\nwas compatible with the value given above and found no\nsignificant PL polarization, which is consistent with the\naboveassumptionthatthenon-resonantprocessesshould\nnot contribute to the spin density[ 21].\nD. Influence of the laser beam cross section\nFigure10shows a representative series of Hanle curves\nmeasured for different pumping times Tpumpin an exter-\nnal fieldBext= 0.3T with a laser power of PL= 39mW.\nSimilar to the data shown above, they show a buildup\nof nuclear spin polarization, manifested as a shift of the\nmaximum towards higher fields. Compared to the data\nshownabove,the monochromatorwassettothe PLmax-\nimum of the d= 39.1nm quantum well, and the optical\ndetuning of the laser was ∆ λ= 7nm, which resulted in\ncorrespondingly lower PL polarization. We used smaller\nincrements for Tpumpin order to achieve higher temporal\nresolution of the pumping process. Using eq.( 6), we fit-\ntedeachcurveseparatelytothe theoreticalshape, adding\nan offset of 0 .3%, which may be due to stray light. As\nshown in fig.( 10), we find the expected increase of the\nnuclear field with the pumping time. In addition, the\ncurves broaden and the maximum of the polarization de-\ncreases with increasing pumping time. To understand\nthese additional effects, we consider the intensity distri-\nbution over the laser beam cross section, which we as-\nsume to be Gaussian. As a result of this distribution,\ndifferent sample regions experience different intensities\nresulting in different pumping rates. To describe this\neffect, we write the laser intensity as a function of the\nFIG. 10. Shifted Hanle curves: measured data and calculated\nHanle curves based on the fit-parameters shown in fig. 12.\ndistance rfrom the center of the beam as\nI(r) =I0·e−r\n2σ2\n0, (14)\nwith the maximum intensity\nI0=PL\n2πσ2\n0. (15)\nFor a laser power PL= 39mW and beam width\nσ0= 88µm, we obtain I0= 0.8µW\nµm2. For a numeri-\ncal simulation of the observed Hanle curves, we calcu-\nlated the pumping dynamics and resulting Hanle curves\nh(Bext,I(r)) for annular segments of the laser beam us-\ning eq.(6). The individual curves were then weighted\nwith the PL power I(r)rdremitted by each ring. The\nresulting calculated Hanle curve is\nSC=´\ndrh(Bext,I(r))rI(r)´\ndrrI(r). (16)\nFigure11shows the calculated Hanle curves SCfor in-\ncreasing times Tpump= [0...2000]s. The saturation of\nthe nuclear field Bnucas well as the broadening of the\ncurves and the decrease in the maximal degree of PL\npolarization are clearly visible. Figure 12compares the\npredictions from this simple model with the experimen-\ntal Hanle curves. The decreasing maximal degree of PL\npolarization and the broadening of the curves, repre-\nsented by ∆ B, as well as the shift of the maxima, |Bnuc|,\nare qualitatively well described by the theoretical model.\nFigure13summarizes this model in a different way: it\nshows the variation of the effective field Bext+Bnucover\nthe laser beam cross section. The dependence of the ef-\nfective field on the position ris calculated for two dif-\nferent pumping times, Tpump= 36s (solid red line) and\nTpump= 271s (dashed red line) with an external field of\nBext= 0.5T, and a laser power of PL= 49mW.7\nFIG. 11. Calculated Hanle curves SCfor increasing times\nTpump= [0...2000]s.\nFIG. 12. Comparison between the experimental parameters\n|Bnuc|, ∆BandS0obtained byfittingthe Hanle curvesshown\nin fig.10(a) using eq.( 6) with the result of the simulation.\nV. DISCUSSION AND CONCLUSION\nAs we have shown, our experimental setup allows one\nto measure the optical pumping dynamics of nuclear\nspins in GaAs quantum wells. We presented a simple\ntheoretical model that describes the buildup of the nu-\nclear spin polarization and therefore of the nuclear field\nin a quantitative way and agrees with the experimental\ndata, within the experimental uncertainties. This model\nstarts with the generation of electron spins by the optical\npumping process. The relevant rate of electron spin den-\nsity production is proportional to the laser power PLand\ndecreases with increasing optical detuning ∆ λ. The spin\npolarizationis then transferredfrom the electron spins to\nthe nuclear spins, and we also determined the rate con-\nstant for this process. The experimental data show some\nFIG. 13. Laser intensity and effective field Bext+Bnucwith\nBext= 0.5T as a function of the position in the laser beam\nin units of σ0= 88µm. The solid red line is calculated for\na pumping time, Tpump= 36s and the dashed red line for\nTpump= 271s.\nsignificantdeviationsfromthe simple model, whichcould\nbe explained quantitatively by taking into account that\nthe laser beam does not illuminate the sample homo-\ngeneously, but with a roughly Gaussian profile. These\nresults can be used to prepare the nuclear spin system\ne.g. for optically detected nuclear magnetic resonance\nexperiments.\nACKNOWLEDGMENTS\nWe gratefully acknowledge the support by the In-\nternational Collaborative Research Centre TRR 160\n“Coherent manipulation of interacting spin excitations\nin tailored semiconductors,” funded by the Deutsche\nForschungsgemeinschaft.8\n[1] S. Pfalz, R. Winkler, T. Nowitzki, D. Reuter, A. D.\nWieck, D. H¨ agele, and M. Oestreich, Phys. Rev. B 71,\n165305 (2005).\n[2] M. DyakonovandV. Perel, in Optical Orientation , edited\nby F. Meier and B. Zakharchenya (Elsevier, 1984), vol. 8\nofModern Problems in Condensed Matter Sciences , pp.\n11 – 71.\n[3] A. Ekimov and V. Safarov, Soviet Journal of Experimen-\ntal and Theoretical Physics 12, 1 (1970).\n[4] R. R. Parsons, Phys. Rev. Lett. 23, 1152 (1969).\n[5] R. I. Dzhioev, B. P. Zakharchenya, V. L. Korenev, P. E.\nPak, D. A. Vinokurov, O. V. Kovalenkov, and I. S.\nTarasov, Physics of the Solid State 40, 1587 (1998).\n[6] S. E. Barrett, R. Tycko, L. N. Pfeiffer, and K. W. West,\nPhys. Rev. Lett. 72, 1368 (1994).\n[7] T. Pietraß, A. Bifone, T. R˜ o˜ om, and E. L. Hahn, Phys.\nRev. B53, 4428 (1996).\n[8] A. I. Ekimov and V. I. Safarov, Soviet Journal of Exper-\nimental and Theoretical Physics Letters 15, 179 (1972).\n[9] O. S. Leifson and C. D. Jeffries, Phys. Rev. 122, 1781\n(1961).\n[10] G. Lampel, Phys. Rev. Lett. 20, 491 (1968).\n[11] A. Abragam, The Principles of Nuclear Magnetism (Ox-\nford University Press, 1967).\n[12] D.Paget, G.Lampel, B.Sapoval, andV.I.Safarov, Phys.\nRev. B15, 5780 (1977).\n[13] G. Salis, D. D. Awschalom, Y. Ohno, andH. Ohno, Phys.\nRev. B64, 195304 (2001).\n[14] M.J. Snelling, G.P. Flinn, A.S. Plaut, R.T. Harley,\nA.C. Tropper, R. Eccleston, and C.C. Phillips, Physical\nReview B 44, 11345 (1991).\n[15] V. Kalevich, V. Korenev, and O. Fedorova, Soviet Jour-\nnal of Experimental and Theoretical Physics Letters 52,349 (1990).\n[16] M. Dyakonov and V. Perel, Soviet Journal of Experimen-\ntal and Theoretical Physics 38, 177 (1974).\n[17] M. Dyakonov, V. Perel, V. Berkovits, and V. Safarov,\nSoviet Journal of Experimental and Theoretical Physics\n40, 950 (1975).\n[18] B. Cavenett, Advances in Physics 30, 475 (1981).\n[19] M. Eickhoff, B. Lenzmann, G. Flinn, and D. Suter, Phys.\nRev. B65, 125301 (2002).\n[20] M. Eickhoff, B. Lenzmann, D. Suter, S. E. Hayes, and\nA. D. Wieck, Phys. Rev. B 67, 085308 (2003).\n[21] M. Eickhoff and D. Suter, J. Magn. Reson. 166, 69\n(2004).\n[22] M. Eickhoff, S. Fustmann, and D. Suter, Phys. Rev. B\n71, 195332 (2005).\n[23] C. Weisbuch and C. Hermann, Phys. Rev. B 15, 816\n(1977).\n[24] R. Hannak, M. Oestreich, A. Heberle, W. R¨ uhle, and\nK. K¨ ohler, Solid State Communications 93, 313 (1995).\n[25] M. J. Snelling, E. Blackwood, C. J. McDonagh, R. T.\nHarley, and C. T. B. Foxon, Phys. Rev. B 45, 3922\n(1992).\n[26] J. Shi-Rong, X. Zhong-ying, L. Jin-sheng, L. Chang-pin g,\nX. Ji-zong, and Z. Bao-zhen, Acta Physica Sinica (Over-\nseas Edition) 3, 384 (1994).\n[27] S. Eshlaghi, W. Worthoff, A. D. Wieck, and D. Suter,\nPhys. Rev. B 77, 245317 (2008).\n[28] W. Hanle, Zeitschrift fuer Physik 30, 93 (1924).\n[29] A. Berg, M. Dobers, R. R. Gerhardts, and K. v. Klitzing,\nPhys. Rev. Lett. 64, 2563 (1990)." }, { "title": "1003.1812v1.Laser_catalyzed_spin_exchange_process_in_a_Bose_Einstein_condensate.pdf", "content": "arXiv:1003.1812v1 [quant-ph] 9 Mar 2010Laser-catalyzed spin-exchange process in a Bose-Einstein condensate\nH. Jing1,2, Y. Jiang1, Weiping Zhang2, and P. Meystre3\n1Department of Physics, Henan Normal University, Xinxiang 4 53007, People’s Republic of China\n2State Key Laboratory of Precision Spectroscopy, Departmen t of Physics,\nEast China Normal University, Shanghai 200062, People’s Re public of China\n3B2 Institute and Department of Physics, The University of Ar izona, Tucson, Arizona 85721, USA\n(Dated: November 5, 2018)\nWe show theoretically that it is possible to optically contr ol collective spin-exchange processes\nin spinor Bose condensates through virtual photoassociati on. The interplay between optically in-\nduced spin exchange and spin-dependent collisions provide s a flexible tool for the control of atomic\nspin dynamics, including enhanced or inhibited quantum spi n oscillations, the optically-induced\nferromagnetic-to-antiferromagnetic transition, and coh erent matter-wave spin conversion.\nPACS numbers: 42.50.-p, 03.75.Pp, 03.70.+k\nSpinor Bose-Einstein condensates, consisting of atoms\nwith internal spin states, provide a promising bridge\nbetween atomic, molecular and optical physics (AMO),\nmatter-wave optics, many-body physics, and quantum\ninformation science [1]. Beside the study of their rich\nground-sate properties [2] and spatial spin structures [3–\n5], the magneticcontrolofquantum spin mixinghasbeen\na central topic of investigations for atomic spin systems\n[6–13]. Also, by steering the spin degrees of freedom\nin ultracold quantum gases, the creation of topological\nskyrmions, Dirac monopoles, dark solitons, and quantum\nentanglement have been investigated [14–16].\nIn parallel to these works, which concentrate largely\non the role of external magnetic fields on spinor con-\ndensates, there have also been important developments\non their magneto-optical manipulation. For example,\nZhangetal.[17] studied the spin waves induced by light-\ninduceddipole-dipoleinteractionsinanatomicspinchain\ntrapped in a lattice potential. More recently, advances\nin coherent photoassociation (PA) at ultracold tempera-\ntures [18] were exploited in theoretical and experimental\nstudies of spin mixing and spin-dependent PA in spin-1\ncondensates [19, 20]. In very recent work, Kobayashi\net al.observed the spin-selective formation of spinor\nmolecules in ferromagnetic atoms87Rb [21].\nIn this Rapid Communication, we demonstrate theo-\nreticallytheopticalcontrolofatomicspinmixingin afer-\nromagnetic spin-1 Bose gas. In the proposed method the\nopticalfieldsinducevirtualPAprocessesintheatoms,for\nexample via a dark molecular state. We show that this\nstep, which can be intuitively coined a laser-catalyzed\nspin-exchange process (LCSE), opens up an efficient and\nwell-controlled optical channel for coherent atomic spin\nmixing. By tuning the the strength ratio of these two\nchannels of LCSE and spin-dependent collisions, three\ndifferent regimes can be identified in the laser-controlled\nquantum spin dynamics, i.e., going from the collision-\ndominated regime to the no-spin-mixing regime, and to\nthelaser-inducedantiferromagneticregime,whichisrem-\niniscent of the prominent role of long-range dipole-dipole\ninteraction in a dipolar spin gas (by changing the ratio of\nspin-dependent collision and dipolar interaction). As aresult, and in contrast to the collision-dominated ”single-\nchannel” case, a wealth of important new effects arise in\nthe spin dynamics of the atoms.\nFor concreteness we concentrate on two limiting situ-\nations, the adiabatic far off-resonant regime and the res-\nonant case [22]. Our purpose here is to show that the\ninterplay of two channels, the spin-dependent collisions\nand LCSE, provides a flexible tool for the control of the\natomic spin dynamics, including enhanced and inhibited\nquantum spin oscillations, laser-induced ferromagnetic-\nto-antiferromagnetic (F-AF) transitions, as well as effi-\ncient coherent spin transfer triggered even by quantum\nvacuum noise. This method can be also extended to\nstudy e.g. the optical control of domain formation and of\nspin textures in a spinor gas [4]. As such, optical LCSE-\ncontrolled spinor condensates provide a promising new\ntool for the study of collective chemically-driven quan-\ntum spin dynamics.\nFigure 1 illustrates the process under consideration,\nthe dynamics of a spin-1 atomic condensate resulting\nfrom the virtual PA of two mF= 0 atoms into a molecu-\nlar state, followed by dissociation into a pair of mF=−1\nandmF= +1 atoms. Accounting in addition for spin-\ndependent collisions between atoms, this system is de-\nscribed by the Hamiltonian ( /planckover2pi1= 1)\nH=Hcoll+Hpa, (1)\nwhereHcollandHparefer to spin-dependent collisions\nand the controllable light-assisted interactions, respec-\ntively, with\nHcoll=/integraldisplay\ndr/bracketleftBig\nˆψ†\ni(−/planckover2pi12\n2m∇2+V+Ei)ˆψi+c′\n0\n2ˆψ†\niˆψ†\njˆψjˆψi\n+c′\n2\n2ˆψ†\nkˆψ†\ni(Fγ)ij(Fγ)klˆψjˆψl/bracketrightBig\n, (2)\nHpa=/integraldisplay\ndr/bracketleftBig\n∆′ˆψ†\nmˆψm+Ωp(ˆψ†2\n0ˆψm+ˆψ†\nmˆψ2\n0)\n−Ωd(ˆψ†\nmˆψ+ˆψ−+ˆψ†\n−ˆψ†\n+ˆψm)/bracketrightBig\n. (3)\nHereˆψi(r) is the annihilation operatorof the i-th compo-\nnent atom ( i= 0,±1) andˆψm(r) the corresponding op-2\nFIG. 1: (Color online) Schematic of coherent two-channel\nspin-exchange interactions in an optically-controlled sp in-1\nBose gas. In addition to the familiar collisional channel, t he\noptical channel proceeds via the virtual PA of two |mF= 0/angbracketright\natoms into an intermediate molecular state |m/angbracketright, which disso-\nciates into a pair of |mF= +/angbracketrightand|mF=−/angbracketrightatoms.\nerator for the intermediate molecular field, Vis the trap\npotential,Eiis the Zeeman shift and Fγ=x,y,zis the spin-\n1 matrix [3]. The coefficients c′\n0= 4π/planckover2pi12(a0+ 2a2)/3m\nandc′\n2= 4π/planckover2pi12(a2−a0)/3mgive the strength of the spin-\nconservingtwo-body collisions, with a0,a2the scattering\nlengths of the accessible collision channels, and mis the\natomic mass. The Rabi frequencies Ω p,ddescribe the\nstrength of the photoassociation of mF= 0 atoms and\ndissociation into mF=±1 atoms, and ∆′is the detuning\nbetween the molecular and atomic states.\nWe first consider the adiabatic off-resonant regime by\nassuming that ∆′is the largest parameter in the system.\nFrom∂ˆψm/∆′∂t≃0, we have ˆψm≃Ωdˆψ+ˆψ−−Ωpˆψ2\n0\n∆′, and\nsubstitute this form into the Heisenberg equations of mo-\ntion derived from Eq. (1) [23]. It is easily seen that the\nresulting equations can also be derived from the effective\nHamiltonian\nHeff=Hcoll+/integraldisplay\ndr/bracketleftBig\nΩ′(ˆψ†\n+ˆψ†\n−ˆψ2\n0+ˆψ†2\n0ˆψ+ˆψ−)+O/bracketrightBig\n,(4)\nwith\nΩ′= ΩpΩd/∆′,O=−Ω2\np\n∆′ˆψ†2\n0ˆψ2\n0−Ω2\nd\n∆′ˆψ†\n−ˆψ†\n+ˆψ+ˆψ−.\nIn addition to the familiar spin-dependent collisions, spin\ncoupling now also results from optical LCSE, resulting in\nthe two-channel spin-exchange Hamiltonian\nH=C′/integraldisplay\ndr(ˆψ†\n+(r,t)ˆψ†\n−(r,t)ˆψ0(r,t)ˆψ0(r,t)+h.c.).\nHereC′= Ω′+c′\n2describes the combined effect of spin-\ndependent collisionsand optical LCSE.Clearly, the effec-\ntive spin-coupling strength C′can be negative or positivewith suitable optical parameters, leading to significantly\ndifferent spin dynamics.\nIn the limit where the spatial degrees of freedom de-\ncouple from the spinor dynamics, that is, when the spin\nhealing length is larger than the condensate size, it is\npossible to invoke the single-mode approximation [9, 13]\nwhereˆψi(r,t)→φ(r)ˆai(t),ˆψm(r,t)→φ(r)ˆm(t), where\nφ(r) is the spatial wave function of the condensate with\nˆaiand ˆmbeing the atomic or molecular annihilation\noperators. The rest of this paper presents results ob-\ntained in this approximation. For convenience we also\nintroduce the scaled parameters c0,2=c′\n0,2/integraltext\ndr|φ(r)|4,\nΩ = Ω′/integraltext\ndr|φ(r)|4, ∆ =Ω2\nd\n∆′/integraltext\ndr|φ(r)|4, andτ=c0nt,\nwherenis the initial atomic density.\nThe mean-field evolution of the spin-0 atomic popu-\nlationn0is illustrated in Fig. 2 for several values of\nC= Ω +c2and for the initial state ( n+, n0, n−) =\n(0.05,0.9,0.05). The specific example of atoms87Rb\nis considered here, with a0= (101.8±2)aBanda2=\n(100.4±1)aB[24], where aBis the Bohr radius. In our\ncalculations, we assume the optical detuning ∆′= 100Ωp\nand Ωd= 10Ωp, which is reasonable for the adiabatic\noff-resonant case. The typical initial atomic density\nn∼1014cm−3,corresponding to c0n∼9700Hz.\nSignificantly different regimes are reached by varying\nthe optical detuning ∆′: (i) In the collision-dominated\nregime|c2|>|Ω|, the population of spin-0 atoms is al-\nwayslargerthan its initial value 0 .9. In this perturbed\nregime, the spin coupling is still ferromagnetic ( C<0);\n(ii) ForC= 0, i.e. Ω = −c2(for ∆′>0) the two channels\n(spin-dependent collisions and LCSE) interfere destruc-\ntively, leadingtofrozenorinhibitedspinmixing; (iii)The\nsignofCcan be reversedby tuning the laserfields, result-\ning in an effective antiferromagnetic regime C>0. This\nis reminiscent of the F-AF transition induced by long-\nrange dipolar interactions in a spinor gas [10]. In this\nreversed regime, the atomic spin oscillations ( |Ω|>|c2|)\nturn out to be belowthe initial value 0 .9.\nThe mean-field atomic spinor dynamics is well de-\nscribed by a nonrigid pendulum model [3]. By expressing\nthec-number amplitudes aiin terms of real amplitudes\nand phases, i.e. a±,0=√n±,0e−iθ±,0, the energy func-\ntional of the optically-controlled spinor system can be\nderived as\nE=q(1−n0)+C/radicalbig\n(1−n0)2−m2cosθ\n+c2n0(1−n0)+∆\n4n0(2−n0)−Ω2\n∆n2\n0(5)\nwhereqdenotes the quadratic Zeeman effect, θ=θ++\nθ−−2θ0is the relative phase of the spin components and\nm=n+−n−is the atomic magnetization. The last two\nterms in Eq. (5) account for the optical energy shift.\nThe effective spin-coupling parameter Chas an impor-\ntant impact on the properties of the system, as already\nmentioned. Figure 3 plots equal-energy contours in the\nphase space ( θ,n0), for several values of Cform= 0 and\nq= 0.01, corresponding to the fixed magnetic field about3\n460mGin our specific example, which is larger than the\nresonance magnetic field Bres∼330 mG, see [27]. We\nhave also carried out numerical simulations for values of\nqcorrespondingto B 0,x= 1, and y= 1/4−γH. HereγHis proportional\nto Hund’s coupling JH, supporting FM correlations in\nthe spin sector. For S= 1 and realistic values of Hund’s\ncoupling for vanadates, γH∼0.1, numerical investiga-\ntions ofthis model showedstrongbut short-rangeddimer\ncorrelations in a certain finite temperature range caused\nby the related entropy gain although the ground state is\nuniformly FM.24The same Hamiltonian was also studied\nusing a MF decoupling scheme.26Within this approach a\nfinite temperature phasewith dimer orderin both sectors\nwasfound. However,asdiscussedintheintroduction, the\nMF decoupling approach has severe limitations as it does\nnot take the coupled spin-orbital dynamics into account.\nWe start from a generalizationof Eq. ( 1.1) which reads\nHSτ(Γ) =J/summationdisplay\nj(Sj·Sj+1+x)/parenleftbig\n[τj·τj+1]Γ+y/parenrightbig\n,(2.1)\nwith\n[τj·τj+1]Γ≡τj·τj+1−Γτz\njτz\nj+1.(2.2)\nThe calculations presented here are valid for general S\nandx, but we will, unless stated otherwise, only address4\n-1.5-1-0.500.51\n0 1\nT/J-0.4-0.200.20.4<τx\niτx\ni+1+τy\niτy\ni+1>y=0.4y=-0.3\ny=0.4\ny=-0.3\nFIG. 2: (Color online) Nearest-neighbor spin and orbital co r-\nrelation functions for the spin-orbital model (2.1) with Γ = 1\nas a function of temperature Tin units of J(we setkB= 1).\nIn both panels y= 0.4, 0.3, 0.2, 0.15, 0.1, 0.05, 0.0, −0.1,\n−0.2,−0.3 in arrow direction. The spin correlations switch\nfrom AF to FM at y= 0.1 (dashed lines). The dotted lines\nin the upper (lower) panel correspond to the limiting values\n1 and−1.4015 (−1/πand 1/π), respectively.\nthe caseS= 1,x= 1 relevant for YVO 3in the following.\nFor Γ = 1 the pseudospin sectorreduces to an XY model.\nIn Fig.2the nearest-neigbor spin and orbital corre-\nlation functions for Γ = 1 as a function of temperature\nfor various parameters yobtained by TMRG are shown.\nThis method allows us to obtain thermodynamic quan-\ntities for 1D quantum systems directly in the thermody-\nnamic limit.47–49Fory/lessorsimilar0.1 the ground state has ferro-\nmagneticallyalignedspins. Thephasetransitionbetween\nthefullypolarizedFMstateandastatewithAFspincor-\nrelations at y≈0.1 is first order. In the limit y≫1 the\nvalue/an}bracketle{tSjSj+1/an}bracketri}ht ≃ −1.4015 for a Haldane S= 1 Heisen-\nberg chain is reached,50while the orbital correlations ap-\nproach/an}bracketle{tτx\njτx\nj+1+τy\njτy\nj+1/an}bracketri}ht →1/π. We note that the FM\nground state is lost at y≃0.1 both in the model with\nΓ = 1 investigated here as well as in the model with an\nisotropic pseudospin sector (Γ = 0).24,51For the model\nwith Γ = 1, however, we have a direct phase transition\nfrom the FM to the Haldane phase while for the isotropic\nmodel an orbital valence bond phase is intervening be-\ntween these two phases.51\nThe isotropic model ( 2.1), Γ = 0, has also been in-\ntensely studied for S=τ= 1/2. Here the phase diagram\nis more complex than in the S= 1,τ= 1/2 case.52,53Forx= 1 the FM spin state is again found to be stable\nfory/lessorsimilar0.1. However, now the transition at y∼0.1 is to\na gapless “renormalized SU(4)” phase followed by a fur-\nther phase transition at larger yinto a dimer phase. The\nphase with ferromagnetically polarized orbitals is absent\nbecause/an}bracketle{tSj·Sj+1+x/an}bracketri}ht>0 forx= 1 but is again present\nfor large yifx/lessorsimilarln2−1/4.\nB. Boson-fermion model\nFor the 1D spin-orbital model, Eq. ( 2.1), at the point\nΓ = 1, we will now derive an effective BF model which\nwill be used as a starting point for the perturbative\napproach. First, applying the Jordan-Wigner transfor-\nmation the orbital part is mapped onto a free fermion\nmodel (for Γ /ne}ationslash= 1 the pseudospins map onto interacting\nfermions). The spin part of the spin-orbital Hamiltonian\n(2.1) will be represented by bosons. Concentrating on\nthe case where the spin part is ferromagnetically polar-\nized in the ground state, we can treat the spin sector by\nthe MSWT.41,42To this end, we introduce bosonic op-\nerators by a Dyson-Maleev transformation. If we retain\nbosonic operators only up to quadratic order we end up\nwith\nH≡ HSτ(1)−JN(S2+x)y≃H0+H1.(2.3)\nHereH0is already diagonal\nH0=/summationdisplay\nkωB(k)b†\nkbk+/summationdisplay\nqωF(q)f†\nqfq,(2.4)\nwithf†\nqandfq(b†\nkandbk) being the fermionic (bosonic)\ncreation and annihilation operators, respectively. The\nmagnon dispersion is given by\nωB(k) = 2JS|y|(1−cosk), (2.5)\nand the fermion dispersion reads\nωF(q) =J(S2+x)cosq. (2.6)\nThe spinless fermions fill up the Fermi sea between the\nFermi points at kF=±π/2.\nFor FM spin chains usual spin-wave theory has to be\nmodified by a Lagrange multiplier µacting as a chemical\npotential which enforces the Mermin-Wagner theorem of\nvanishing magnetization at finite temperature41\nS=1\nN/summationdisplay\nk/angbracketleftBig\nb†\nkbk/angbracketrightBig\n. (2.7)\nThermodynamic quantities calculated with this method\nare in excellent agreement with the exact Bethe ansatz\nsolution for the uniform chain as well as with numerical\nTMRGdataforthedimerizedFMchainfortemperatures\nup toT∼ |Jeff|S2, withJeffbeing the effective exchange\nconstant of the model under consideration.26,41,42,545\nThe interacting part couples bosons and fermions and\nreads\nH1=1\nN/summationdisplay\nk1,k2,qωBF(k1,k2,q)b†\nk1bk2f†\nqfk1−k2+q,(2.8)\nwith the vertex\nωBF(k1,k2,q)≡JS[cos(k2−q)+cos(k1+q)\n−cos(k1−k2+q)−cosq].(2.9)\nThe Hamiltonian H=H0+H1withH0andH1given\nby Eqs. ( 2.4) and (2.8), supplemented by the constraint\n(2.7), is an effective BF representation valid at low tem-\nperatures. We will investigate this model in Sec. Vtreat-\ning the BF coupling perturbatively.\nIII. MEAN-FIELD DECOUPLING\nThe spin-orbital model ( 2.1) contains rich and inter-\nesting physics. A first attempt to understand the prop-\nerties of the model is to apply a MF decoupling which\nneglects the coupled spin-orbital degrees of freedom and\ntreats the spin-orbital chain as two separate chains with\neffective coupling constants which have to be determined\nself-consistently. Note, however,that this treatment does\nnot involve site variables as in the classical Weiss-MF\ntheory but takes the correlations on a bond as relevant\nvariables. Interestingly, these expectation values never\nvanish, which makes them useful particularly in cases\nwithout long-range order.\nA. Decoupling into spin and orbital chain\nApplying a MF decoupling and allowing for a dimer-\nization in both sectors26,27we obtain from Eq. ( 2.1)\nHSτ(Γ)≃HMF\nS+HMF\nτ(Γ), (3.1)\nwith the spin and orbital Hamiltonians\nHMF\nS=JSN/summationdisplay\nj=1{1+(−1)jδS}Sj·Sj+1,\nHMF\nτ(Γ) =JτN/summationdisplay\nj=1{1+(−1)jδτ}[τj·τj+1]Γ.(3.2)\nWithin this approximation the effective superexchange\nconstants and dimerization parameters are given by\nJτ=J∆+\nSS+2x\n2, δ τ=∆−\nSS\n∆+\nSS+2x,\nJS=J∆+\nττ+2y\n2, δ S=∆−\nττ\n∆+ττ+2y,(3.3)where we have defined\n∆±\nSS=/an}bracketle{tS2j·S2j+1/an}bracketri}ht±/an}bracketle{tS2j·S2j−1/an}bracketri}ht,\n∆±\nττ=/angbracketleftbig\n[τ2j·τ2j+1]Γ/angbracketrightbig\n±/angbracketleftbig\n[τ2j·τ2j−1]Γ/angbracketrightbig\n.(3.4)\nHere ∆−\nσσwithσ=S(σ=τ) is an order parameter for\nthe spin (orbital) dimerization, respectively. Thus, the\nexchangeconstantsand dimerization parametersfor each\nsectoraredeterminedbythe nearest-neighborcorrelation\nfunctions in the other sector, making a self-consistent\ncalculation necessary. In the following we want to solve\nEqs. (3.2)-(3.4) for the spin exchange being effectively\nFM, i.e. JS<0.\nB. Dimerized orbital correlations\nNumerical investigations of the model with isotropic\norbital exchange, HSτ(0), have shown orbital-singlet for-\nmation in the ground state51fory/greaterorsimilar0.1. Moreover,\nalthough the ground state consists of a fully spin polar-\nized FM state for y/lessorsimilar0.1 with AF orbital correlations,\nit has been shown that a tendency towards orbital sin-\nglet formation is still present but has to be activated by\nthermal fluctuations.24In Ref. 26 the model ( 3.1) was\nstudied in the FM regime with x= 1,y=1\n4−γH,\nΓ = 0 and γH= 0.1 in order to address the ques-\ntion whether this orbital-Peierls effect can be captured\nwithin a MF decoupling approach. A dimerized phase\nfor 0.10/lessorsimilarT/J/lessorsimilar0.49 (we set kB=/planckover2pi1= 1) was found\nwith the dimerization amplitude in the spin sector being\nmuch larger than in the orbital sector.\nWe now want to compare this result with the case\nwhere we set Γ = 1 in Eq. ( 3.1) so that the self-\nconsistent Eqs. ( 3.3) can be solved analytically by ap-\nplying a Jordan-Wigner transformation and MSWT. In-\ntroducing fermionic operators f(†)\nj,eif the index jis even\nandf(†)\nj,oifjis odd for the pseudospins, we rewrite\nHMF\nτ≡HMF\nτ(Γ = 1) in Fourier representation. Finally\nintroducing new fermionic operators φ(†)\nqandϕ(†)\nqwhich\ndiagonalize the Hamiltonian HMF\nτ, we find\nHMF\nτ=/summationdisplay\nqωMF\nF(q,δτ)(φ†\nqφq+ϕ†\nqϕq),(3.5)\nwith the fermionic dispersion55\nωMF\nF(q,δτ)≡ Jτ/radicalBig\ncos2q+δ2τsin2q.(3.6)\nWe can now calculate ∆±\nττ, as given in Eq. ( 3.4)\nstraightforwardly and obtain\n∆−\nττ=2δτ\nN/summationdisplay\nq/braceleftbig\n2nF[ωMF\nF(q,δτ)]−1/bracerightbig\nsin2q/radicalBig\ncos2q+δ2τsin2q,\n∆+\nττ=2\nN/summationdisplay\nq/braceleftbig\n2nF[ωMF\nF(q,δτ)]−1/bracerightbig\ncos2q/radicalBig\ncos2q+δ2τsin2q,(3.7)6\nwherenF(x) ={exp(βx)+1}−1is the Fermi function\nandβ= 1/T.\nC. Dimerized spin correlations\nNext we turn to the spin part of Eq. ( 3.1) to which\nwe apply the MSWT.41,42We introduce two bosonic op-\neratorsb(†)\nj,e[b(†)\nj,o] forjeven [odd] by means of a Dyson-\nMaleev transformation. Retaining only terms bilinear in\nthe bosonic operators we can diagonalize the resulting\nHamiltonian by a Bogoliubov transformation leading to\nHMF\nS=/summationdisplay\nk/braceleftBig\nωMF\nB,−(k,δS)α†\nkαk+ωMF\nB,+(k,δS)β†\nkβk/bracerightBig\n+JSNS2,\n(3.8)\nwith the two magnon branches\nωMF\nB,±(k,δS) = 2|JS|S/parenleftbigg\n1±/radicalBig\ncos2k+δ2\nSsin2k/parenrightbigg\n.(3.9)\nThe constraint of vanishing magnetization at finite tem-\nperature ( 2.7) now reads\nS=1\nN/summationdisplay\nk/braceleftbig\nnB[ζ−(k,δS)]+nB[ζ+(k,δS)]/bracerightbig\n,(3.10)\nwherenB(x) ={exp{βx}−1}−1is the Bose function and\nζ±(k,δS) =ωMF\nB,±(k,δS)−µ(δS).\nTo calculate the nearest-neighborcorrelationfunctions\nB±≡/angbracketleftbig\nSj·Sj±1/angbracketrightbig\nit is necessaryto go beyond linear spin-\nwave theory. Taking terms of quartic order into account\nand using Eq. ( 3.10) we obtain26\nB±=\n1\nN/summationdisplay\nkf±(k,δS)/summationdisplay\nσ∈{±}σnB[ζσ\nB(k,δS)]\n2\n.\n(3.11)\nHere we have defined\nf±(k,δS)≡cos2k±δSsin2k/radicalBig\ncos2k+δ2\nSsin2k.(3.12)\nFrom these expressions we can obtain ∆±\nSSwhich, com-\nbined with Eq. ( 3.7), allows us to solve Eqs. ( 3.1)-(3.4)\nself-consistently.\nD. Mean-field phase diagram\nWe first discuss the ground state phase diagram of the\nHamiltonian ( 3.1) for Γ = 1. Depending on the sign of\nthe effective coupling constant JSwe find/an}bracketle{tSjSj+1/an}bracketri}ht=\n1,−1.4015 with the latter value being the approximate\nresult for the S= 1 AF Haldane chain. In the following,\nwe restrict our discussion to −1< x <1.4015 so that0.1 0.15 0.2y00.10.20.30.4 T/Jdimerizeduniform\nuniform(a)\nytpT2T1\ntricritical \npoint\n0.2 0.3 0.4\nT/J00.51Dimerization parameters(b)δS\nδτ\nT2/J T1/Jdimerized\nFIG. 3: (a) Phase diagram of the Hamiltonian (3.1) with\nΓ = 1 and x= 1 in mean-field decoupling. The shaded area\nrepresents the dimerized phase. The phase transition at T2\nis first order whereas the transition at T1is of second order.\nThe two transition lines merge at the tricritical point ytp. (b)\nDimerization parameters δSandδτforx= 1 and y= 0.14.\nThe lines are guides to the eye. The shaded area marks the\ntemperature range where the dimerization is nonzero.\n/an}bracketle{tSjSj+1/an}bracketri}htandJτ=J(/an}bracketle{tSjSj+1/an}bracketri}ht+x) always have the\nsame sign. For the orbital sector we obtain, on the other\nhand,/an}bracketle{tτx\njτx\nj+1+τy\njτy\nj+1/an}bracketri}ht=±1/π.y >1/πimpliesJS>0\nand the ground state is therefore certainly AF (Haldane\nphase) whereas JS<0 fory <−1/πleading to a FM\nstate. In the regime −1/π < y < 1/πthe self-consistent\nequationshave twosolutions with energies EAF\n0≈(1/π+\ny)(−1.4015+x) andEFM\n0= (−1/π+y)(1+x) and a first\norder phase transition between the FM and AF states\noccurs where the energies cross. For the case x= 1 we\nare focussing on here, this happens at yc≈0.212 and the\nFM stateis stablefor y < yc. Comparedto the numerical\nsolution where yc≈0.1 (see Fig. 2) the range of stability\nof the FM state is therefore increased in the MF solution.\nNext, we investigate the possibility of a finite tem-\nperature dimerization for x= 1 in that part of the\nphase diagram where the ground state is FM. As shown7\nin Fig.3(a) we find that a dimerized phase at finite\ntemperatures does indeed exist in MF decoupling for\nytp≈0.128/lessorsimilary/lessorsimilaryc≈0.212 where ytpdenotes the\ntricrictal point. As in the model with an isotropic pseu-\ndospin sector,26the temperature range where the dimer-\nized phase is stable depends on y. At the onset temper-\natureT1the phase transition is of second order whereas\nat the reentrance temperature T2it is of first order, see\nFig.3(b). As in the case Γ = 0, the dimerization in\nthe spin sector is always much larger than in the orbital\nsector.\nAs pointed out before, the MF decoupling suffers from\nsevere limitations and it is expected to be an even worse\napproximation in the extreme quantum case Γ = 1 than\nin the case Γ = 0 studied previously.26In particular the\ncoupling between spin and orbital degrees of freedom is\ncompletely lost within this approach. In the following\nsections we will therefore develop an alternative pertur-\nbative treatment of the spin-orbital coupling.\nIV. DYNAMICAL SPIN STRUCTURE FACTOR\nFOR THE UNIFORM FERROMAGNETIC CHAIN\nIn order to investigate coupled spin-orbital degrees of\nfreedom and, in particular, their implications on the spin\ndynamics of the spin-orbital chain, a detailed under-\nstanding of the spin dynamics of a FM chain is useful.\nWe shall avoid the complications of the dimerized chain\nand focus our study on the uniform 1D ferromagnet.56\nIn doing so we neglect the coupling between spin and\npseudospin operators for a moment and consider\nHS=JS/summationdisplay\njSj·Sj+1, (4.1)\nwithJS<0. It is well-known that MSWT does not\nrespect the SU(2) symmetry of the FM Heisenberg chain\nEq. (4.1). We therefore directly calculate the full spin\ncorrelation function39,40\nG(r,τ)≡ −/an}bracketle{tT[Sj(0)·Sj+r(τ)]/an}bracketri}ht.(4.2)\nIn Fourier space we obtain\nG(q,ων,B)=1\nN/summationdisplay\nk(1+nB[ζ(k)])nB[ζ(q−k)]1−e−βǫq(k)\niων,B−ǫq(k),\n(4.3)\nwhere we have used the bosonic Matsubara frequencies\nων,Bandǫq(k)≡ζ(k)−ζ(q−k) withk∈[−π,π]. The\nreduced magnon dispersion reads\nζ(k) = 2JSS(1−cosk)−µ. (4.4)\nIn Fig.4(a) the dynamical spin structure factor,\nS(q,ω) = 2nB(−ω)ImGret(q,ω),(4.5)0 1 2 3 4\nω/|JS|050100 |JS|S(q,ω)\n3 3.5\nω/|JS|02040|JS|S(q,ω)\n3 3.5 3.8025 4 4048\n|JS|ρ(ω)q=0q=π/5q=2π/5q=3π/5q=4π/5q=π(a)\n(b)\nω4π/5max\nFIG. 4: (Color online) (a) Dynamical spin structure factor\nS(q,ω) as obtained for T/|JS|= 0.1 and 0 ≤q≤π. The\ndashed line indicates the upper boundary of the two magnon\ncontinuum. The dots are projections of the peak positions\nonto the ( q,ω) plane. They are connected by the dotted line\nwhich is a guide to the eye. (b) Dynamical spin structure\nfactorS(q,ω) for the same parameters at q= 4π/5 (solid\nline) and the corresponding density of states (dashed line) .\nis shown for the uniform FM chain at ω >0, where\nImGret(q,ω) =π\nN/summationdisplay\nk(1+nB[ζ(k)])nB[ζ(q−k)]\n×/parenleftBig\ne−βǫq(k)−1/parenrightBig\nδ(ω−ǫq(k))(4.6)\nis the imaginary part of the retarded Green’s function\nobtained from Eq. ( 4.3) by analytical continuation. Up\nto a factor of 2 π, as a matter of definition, we obtain\nthe result previouslygivenbyTakahashi.39The structure\nfactor fulfills detailed balance, S(q,ω) = eβωS(q,−ω).8\nThe symbols in Fig. 4show the peak positions projected\nonto the ( q,ω) plane. They follow the reduced disper-\nsion Eq. ( 4.4). Also shown in Fig. 4(a) as a dashed\ncurve is ωmax\nq= 4|JS|Ssinq\n2corresponding to the up-\nper boundary of the two magnon continuum ǫq(k) above\nwhichS(q,ω) is zero in this approximation.\nAt the edge of the two magnon continuum S(q,ω) has\na singularity. In Fig. 4(b) the dynamical spin structure\nfactor for the same parameters as used in Fig. 4(a) is\nshown at q= 4π/5 together with the density of states\nwhich is given by ρq(ω) = 1//radicalBig\n(ωmaxq)2−ω2. Right be-\nlow the singularityat ωmax\nqthe density of states to lowest\norder reads ρq(ωmax\nq−δω)∼1/√\nδω, i.e.,S(q,ω) shows\na square root divergence at the upper threshold. If the\nedge singularity and the central peak are well separated\nthen the spectral weight of the edge singularity is much\nsmaller than the spectral weight of the central peak. If,\non the other hand, the edge singularity is close to the\ncentral peak then the shape of the latter is strongly af-\nfected by the occurence of the edge singularity. In this\ncase the edge singularity gives a significant contribution.\nIt is instructive to analyze S(q,ω) in the limit of small\nq. If the edge singularity and the peak of the struc-\nture factor are well separated, the lineshape of the peak\ncan be obtained approximately. To this end, for small\nqbut|JS|S2q/T≫1 we only retain the leading terms\nof Eq. (4.3). Performing a saddle point approximation\nto lowest order we find S(q,ω)∼nB(−ω)(a(q,ω)−\na(q,−ω)), with\na(q,ω)≈2S|JS|Sq\nξ\n(ω−JSSq2)2+/parenleftBig\nJSSq\nξ/parenrightBig2.(4.7)\nThis Lorentzian lineshape is only valid for low temper-\natures. Here ξ≈ |JS|S2/Tis the correlation length in\nthe low-temperaturelimit.39–42Finally, we want to stress\nthat forT→0 the peaks will reduce to δ-functions, i.e.,\nonly thermal broadening is included in this approxima-\ntion.\nV. PERTURBATION THEORY\nIn this section we intent to go beyond the MF decou-\npling approach treating the influence of the BF inter-\naction, Eq ( 2.8), on the spin-wave dispersion perturba-\ntively. Naively one would expect that the magnon should\nbe able to couple to the fermionic degrees of freedom if\nit lies inside the fermionic two-particle continuum. The\nupper and lower boundary of the latter are given by\nǫmax\nF(q) = 2J(S2+x)sin(q/2),\nǫmin\nF(q) =J(S2+x)sinq, (5.1)\nrespectively. The continuum and the magnon dispersion\nωB(q) fory=−1 are shown in Fig. 5. One would there-\nfore expect that in this case ωB(q) is unaffected by the0 0.5 1\nq/π01234 ω/JωB(q)\nFIG. 5: (Color online) Magnon dispersion ωB(q) (dashed line)\nfory= 1 andfermionic two-particle continuum(shaded area).\npresenceofthe fermionsfor q < π/2, sinceit can not cou-\nple to these degrees of freedom. However for higher mo-\nmenta the spin wave may couple to the fermionic degrees\noffreedomandthus abroadeningof S(q,ω)shouldoccur.\nMoreover by choosing different values for ythe point at\nwhich the magnon enters the fermionic two-particle con-\ntinuum is changed. Thus the momentum at which the\nspin wave is affected by the coupling to the fermionic de-\ngrees of freedom depends directly on the parameter y.\nThese argumentsgive the qualitatively correct picture,\ni.e., we find indeed that the coupling of the magnon to\nthe fermionic degrees of freedom has strong effects on\nthe dynamical spin structure factor at intermediate and\nhigh momenta and that the onset of these effects can be\nwell estimated by our simple argument. However, there\narealsocertain aspects which can not be captured within\nthis picture. Forinstance, for 2 S|y|>(S2+x) it suggests\nthat the spin wave may leave the fermionic two-particle\ncontinuum at a certain momentum qland thus should be\nunaffected by the BF coupling for q > ql. The detailed\ncalculation, however, revealsthat this is not true because\nthe spin wave decays into a fermionic particle-hole anda\nremainingspin waveaswill becomeclearinthe following.\nA. General formulation\nHere we want to study the Hamiltonian Hgiven by\nEq. (2.3), with its noninteracting part H0and interact-\ning part H1defined by Eqs. ( 2.4) and (2.8), by treat-\ning the BF interaction perturbatively, i.e., in the limit\n|x|,|y| ≫1. As explained in the appendix, we start by\nperforming a MF decoupling for the interaction H1. Cor-\nrections to this solution are then taken into account per-\nturbatively. Here we adress the bosonic Green’s function\nat zero temperature\nGB(q,t) =−i/angbracketleftbig\nTt/bracketleftbig\nbq(t)b†\nq(0)/bracketrightbig/angbracketrightbig\n. (5.2)9\nq qk1 k1k2\nq qq−k1+k2\nk2k1q\nqq−k1+k2 k2 k1(c) (a) (b)\nFIG. 6: Diagrams which contribute to a renormalization of th e magnon in a perturbation theory: (a,b) Diagrams with momen -\ntum exchange between the magnon and the fermions, and (c) dia gram without momentum exchange. All diagrams are second\norder. Fermionic propagators are shown by solid lines, wher eas bosonic propagators are shown as dashed lines.\nq qq−k1+k2\nk2k1\nFIG. 7: Second order diagram for a system of interacting\nfermions with momentum exchange.\nIn Fig.6all distinct, connected diagrams beyond the MF\ndecoupling up to second order are shown.\nWe calculate the Green’s function from the Dyson\nequation\nGB(q,ω) =1\n/braceleftBig\nG(0)\nB(q,ω)/bracerightBig−1\n−Σ(q,ω),(5.3)\nwith\n/braceleftBig\nG(0)\nB(q,ω)/bracerightBig−1\n=ω−ζ(q), (5.4)\nwhereζ(q) is the reduced magnon dispersion defined\nin Eq. ( 4.4) withJS=J(y−1/π), and the self-\nenergy Σ( q,ω) is approximated by the proper self-energy\nΣ2(q,ω) obtained by summing up the diagrams which\ncan be composed of the diagrams shown in Fig. 6.\nThe diagrams shown in Figs. 6(a) and6(b) are of par-\nticular interest because they are the lowest order dia-\ngrams where bosons and fermions exchange momentum.\nThey describe the part of spin-orbital dynamics which\ncannot be captured within the MF decoupling approach\ndiscussed in section III. The diagram shown in Fig. 6(b)\nhas to be thermally activated, i.e., it does not give any\ncontribution at zero temperature. The same is true for\nthe diagram shown in Fig. 6(c). Thus at T= 0 the only\nsecondorderdiagramwhichcontributestotheselfenergyis the one shown in Fig. 6(a) leading to\nΣ+\nBF(q,ω) =−1\nN2/summationdisplay\nk1,k2ω2\nBF(q,k1,k2)\n×Θ[ωF(q−k1+k2)]Θ[−ωF(k2)]\nω−Ω+q(k1,k2)+i0+,(5.5)\nwhere we have abbreviated57\nΩ±\nq(k1,k2)≡ ±ζ(k1)+ωF(q−k1+k2)−ωF(k2),(5.6)\nwithk1,k2∈[−π,π].\nFor systems of interacting fermions we know that per-\nturbation theory in one dimension often leads to infrared\ndivergencies.58,59Such divergencies occur, for example,\nfor the fermionic analogon of the diagram with momen-\ntum exchange, see Fig. 7. These problems can be over-\ncome by the Dzyaloshinski-Larkin solution or bosoniza-\ntion techniques. For the model considered here, however,\nwe find no divergencies within the considered diagrams.\nOne reason for this behavior is a lack of nesting. While\nfor a fermionic interaction as shown in Fig. 7all the\ndispersions in the denominator of Eq. ( 5.5) are approxi-\nmately linear at low energies here one of the dispersions\nis approximately quadratic so that nesting only occurs\nfor singular points. As a further check, we have evalu-\nated the integrals in Eq. ( 5.5) for a constant vertex at\nsmallqandωand did not find any infrared divergencies.\nFor finite temperatures the Matsubara formalism can\nbe applied straightforwardly. The self-energy Eq. ( 5.5)\nnow reads\nΣ+\nBF(q,ων,B) =−1\nN2/summationdisplay\nk1,k2ω2\nBF(q,k1,k2)\niων,B−Ω+q(k1,k2)\n×N+\nF,B(k1,k2,ων,B,T)NF,F(q,k1,k2,ων,B,T),\n(5.7)\nwhere we have abbreviated\nN±\nF,B(k1,k2,ων,B,T)≡nB[ζ(k1)]±nF[ωF(k2)],\nNF,F(q,k1,k2,ων,B,T)≡nF[ωF(q−k1+k2)]\n−nF[ωF(k2)−ζ(k1)].(5.8)10\n0102030JS(q,ω)\n0 2 4 68\nω/J10-810-4100-ImΣ(q,ω)q=3π/10\nq=5π/10\nq=7π/10\nq=9π/10-1-0.500.51\nq/π051015ω/J(a)\n(b)0204060JS(q,ω)\n0 5 10 15\nω/J10-810-4100-ImΣ(q,ω)q=3π/10\nq=5π/10\nq=7π/10\nq=9π/10-1-0.500.51\nq/π051015ω/J(c)\n(d)\nFIG. 8: (Color online) Perturbative results for the BF model at zero temperature with x= 1. In the left (right) panel y=−1\n(y=−2), respectively. (a),(c) S(q,ω) with the inset showing the region for which ImΣ+\nBF(q,ω) is nonzero as a shaded area\nand the renormalized spin-wave dispersion as a dotted line. WhileS(q,ω) is sharply peaked at low momenta, a significant\nbroadening occurs at higher q. Moreover we find additional structures which, as explained in the text, are due to coupled\nspin-orbital excitations. (b),(d) −ImΣ+\nBF(q,ω) as given in Eq. (5.5) for the corresponding values of qshown in (a) and (c),\nrespectively. The coupled spin-orbital excitations show u p as peaks and edges in ImΣ+\nBF(q,ω) (notice the logarithmic scale).\nNote that at finite temperatures both, the reduced spin-\nwave dispersion ζ(q) as well as the fermionic dispersion\nis renormalized due to the MF decoupling applied to\nEqs. (2.8). The respective expressions are given in the\nappendix, see Eqs. ( A.1) and (A.2).\nAt finite temperatures also the diagram shown in\nFig.6(b) contributes and is given by\nΣ−\nBF(q,ων,B) =−1\nN2/summationdisplay\nk1,k2ω2\nBF(q,k1,k2)\niων,B−Ω−q(k1,k2)\n×(1+N−\nF,B(k1,k2,ων,B,T))NF,F(q,k1,k2,ων,B,T).\n(5.9)B. Dynamical spin structure factor\nBelow we present the results obtained by summing up\nthe diagrams shown in Fig. 6in a Dyson series, but re-\nplacingtheexternallegsbytheSU(2)symmetricfunction\ngiven in Eq. ( 4.3). While the perturbative results can,\nstrictly speaking, only be valid for |x|,|y| ≫1 we extend\nthe results here to more physical values |x|,|y| ∼ O(1)\nwhere we still expect perturbation theory to give at least\na qualitatively correct picture. Numerical results ob-\ntained for the dynamical spin structure factor within this\nperturbative approach are shown in Fig. 8forT= 0 and\ny=−1 andy=−2. In both cases S(q,ω) is sharply\npeaked at small momenta whereas a significant broad-\nening occurs at higher momenta. Note, that within the11\n0 0.2 0.4 0.60.8\nq/π0246810 ω/J\n0.5 0.52 0.54\nq/π22.22.42.6ω/J\ny=-0.75y=-1y=-2\nFIG. 9: (Color online) Renormalized magnon dispersions ωq\nin the FM chain for x= 1 and selected values of y. Inset:\nThe most pronounced Kohn anomalies occur at qnearπ/2.\nMSWTS(q,ω) is always a δ-function for the pure spin\nmodel at T= 0, i.e., the broadening here is solely due to\nthe coupling to orbital excitations.\nBy extracting the central peaks of the dynamical spin\nstructurefactoratvariousmomenta, weobtaintherenor-\nmalized spin-wave dispersion ωqwithin the perturbative\napproach. The result of this is shown in the insets of\nFig.8(a,c) and in more detail in Fig. 9. The magnon dis-\npersion is renormalized and small kinks are visible close\ntoq=π/2, which may be interpreted as Kohn anoma-\nlies (see below). The inset of Fig. 9shows the Kohn\nanomalies with a higher resolution. For itinerant ferro-\nmagnets Kohn anomalies are well-known. Here the inter-\naction between the spins of localized ions is mediated by\nan exchange with the conduction electrons.60–66These\nKohn anomalies can thus be used to gain information\nabout the Fermi surface of the conduction electrons.60–62\nHowever to the best of our knowledge Kohn anomalies in\nthe spin-wavedispersionforinsulatingmaterialshavenot\nbeen adressed so far. As we will show below, the Kohn\nanomalies in our case are caused by coupled spin-orbital\ndegrees of freedom.\nApart from extracting the effective spin-wave disper-\nsion from S(q,ω), we also want to discuss the magnon\nbandwidth (full width at half maximum (FWHM)) Γ q\nof the central peaks. A broadening of the zero temper-\nature peaks occurs whenever the imaginary part of the\nself-energy,\nIm Σ+\nBF(q,ω) =−π\nN2/summationdisplay\nk1,k2ω2\nBF(q,k1,k2)Θ[−ωF(k2)]\n×Θ[ωF(q−k1+k2)]δ(ω−Ω+\nq(k1,k2)),\n(5.10)\nis non-zero at ω= Ω+\nq(k1,k2). The contributions within\nthe sums are now determined by the argument of the δ-\nfunction as well as by the constraints given by the Heav-\niside functions. This procedure, for a given set of param-0 0.2 0.4 0.60.8 1\nq/π00.10.20.30.40.5Γq/Jy=-0.75\ny=-1\ny=-2\nFIG. 10: (Color online) Magnon linewidth Γ q(FWHM of\nS(q,ω)) at zero temperature (data points), as obtained for\nx= 1 and the representative values of yindicated in the plot.\nThe lines are guides to the eye.\netersx,y,andS, effectively yields a region within which\nthe spin wave may scatter on fermion pairs. Henceforth\nwecallthisregiontheBFcontinuum. TheBFcontinuum\nis shown in the insets of Figs. 8(a) and8(c) as shaded ar-\neas. The upper boundary of the BF continuum (solid\nlines), which is periodic with a period of 2 π, is given by\n−4JS(y−1/π) + 2J(S2+x) forq= 0, and decreases\nmonotonously from this value with increasing |q|. The\nlower boundary (dashed lines) is periodic with a period\nofπ.\nTo obtain the FWHM of the structure factor, the mag-\nnitude of the contributions to the sums given in Eq.\n(5.10) are essential. Here not only the ( k1,k2)-region\nwhich contributes to the summation but also the magni-\ntude of the vertex ωBFis of importance. We observe that\nthe vertex is small at small momenta but increases at in-\ntermediate and high momenta. This leads to a strong\nincrease of the magnitude of the imaginary part of the\nself-energyas shownin panels (b) and (d) ofFig. 8. From\nthe insets of Fig. 8it becomes clear that the spin-wave\ndispersion enters the BF continuum depending on y. For\nhigher values of |y|the spin wave enters at lower mo-\nmenta. However, since the vertex gives smaller contribu-\ntions at smaller momenta, the broadening of the central\npeaks of the dynamical spin structure factor turns out\nto be smaller the smaller the momenta are at which the\nspin wave enters the BF continuum. This can be seen in\nFig.10where the magnon linewidth Γ qis shown. The\nonset of a finite Γ qsignals the entrance of the spin-wave\ndispersion into the BF continuum and depending on the\nmomentum at which the entrance occurs the increase of\nΓqis either smooth (entranceat lowmomentum) orsteep\n(entrance at high momentum). In addition, we observe\nthat Γ qhas a maximum at the boundary of the Brillouin\nzone for stronger interactions ( y=−0.75 andy=−1\nin Fig.10respectively), whereas for smaller interactions\nwe observe the maximum at smaller momenta followed12\n0 0.2 0.4 0.60.8 1\nq/π0246810 ω/Jωq\nFIG. 11: (Color online) Effective dispersion of the spin\nwaveωqfory=−1 (red solid line) together with coupled\nspin-orbital excitations deduced from −ImΣ2(q,ω), shown\nby (green) dots within the Bose-Fermi continuum (shadded\narea).\nby a decrease of the FWHM towards the zone boundary\n(y=−2 in Fig. 10).\nInterestingly, the coupling to the orbital degrees of\nfreedom does not only give rise to a featureless broaden-\ning ofS(q,ω) but produces additional structures. These\nadditional structures aremost obviousin Fig. 8(a). From\nFig.8(b) it becomes clear that these structures are dom-\ninated by local extrema as well as edges in the imagi-\nnary part of the self-energy. Eq. ( 5.10) shows that such\nextrema can occur if Ω+\nq(k1,k2), Eq. (5.6), becomes sta-\ntionaryas a function ofthe momenta k1,k2as long as the\nHeaviside functions in Eq. ( 5.10) for these momenta are\nnon-zero. The position of the local maxima in the imag-\ninary part of the self-energy is therefore approximately\ngiven by the values Ω+\nq(k1,k2) at these stationary points.\nThese values correspond to the energy of spin-orbital ex-\ncitations into which the initial spin wave can decay, see\nFig.6(a) and which are stable against small redistribu-\ntions of momenta.\nWe conclude that while we do not have completely\nsharp spin-orbital excitations any more as in the Hamil-\ntonian (1.6) with FM exchange considered in the intro-\nduction, there are still characteristic spin-orbital excita-\ntions of finite width within the spin-orbital continuum.\nAsshowninFig. 11wecanextractthe dispersionofthese\ncharacteristic excitations and find that the coupled spin-\norbital excitations are gapless for the parameters con-\nsidered here. However, as can be clearly seen in Figs.\n8(b) and8(d), the weights of the low-energy excitations\nare orders of magnitude smaller than the excitations lo-\ncated at higher energies. Hence the excitations at high\nenergies give the most dominant contribution to the dy-\nnamical spin structure factor. We therefore expect that\nthese excitations will generate additional entropy in the\ncorresponding temperature range which should show up,0 2 4 6\nω/J0102030JS(q,ω)q=3π/10\nq=5π/10\nq=7π/10\nq=9π/10\nFIG. 12: (Color online) Dynamical spin structure factor\nS(q,ω) calculated perturbatively for the spin-orbital chain at\ntemperature T/J= 0.1 withy=−1.\nfor example, in the specific heat which will be studied in\nthe next section.\nMoreover, we find that these coupled spin-orbtial exci-\ntationsareresponsiblefortheKohnanomaliesmentioned\nabove. We observe that the Kohn anomalies at interme-\ndiate momenta occur when the energy of the spin wave\ncoincides with that of a characteristic spin-orbital exci-\ntation. This is different from the Kohn anomaly in the\nspin-wave dispersion of itinerant ferromagnets. In this\ncase the interaction between the localized spins given by\nthe lattice ions is induced by scattering with conduction\nelectrons and hence the Kohn anomaly is determined by\nthe shape of the Fermi surface.60–66The Kohn anomaly\nwe find within the present context is also due to interac-\ntioneffects, wherethenatureoftheinteractions-coupled\nspin-orbital degrees of freedom - is distinct from the ones\nof the itinerant ferromagnets. For the case x=−y= 1\nthe spin-wave dispersion has a discontinuity of the order\n∆ω≃0.01Jat the point q≃0.509π(see inset of Fig. 9).\nHowever, for the crossing points of the magnon and the\ncoupled spin-orbital excitation located at q≃0.37πand\nq≃0.76π(see Fig. 11) no Kohn-anomaly could be re-\nsolved. We believe that this is a consequence of the\nweight of the coupled spin-orbital excitations: Whereas\natq≃0.509πthe imaginary part of the self-energy dis-\nplays a steep increase of several magnitudes, at the other\ncrossing points the slope towards the local maxima is far\nmore moderate.\nAt finite temperatures two effects contribute to the\nbroadening of the central peaks of the dynamical spin\nstructure factor. First, there is a broadening due to ther-\nmally excited magnons which is already present in the\n1D Heisenberg chain discussed in Sec. IV. This is com-\nbined with the broadening due to the interaction with\nthe orbital degrees of freedom. Here the BF continuum\nis smeared out by thermal fluctuations compared to the\nzero temperature case. Results for the structure factor13\n0 0.2 0.4 0.60.8 1\nq/π00.10.20.30.40.5Γq/JT/J=0.2\nT/J=0.1\nT/J=0\nFIG. 13: (Color online) Magnon linewidth Γ q(FWHM of\nS(q,ω)) as a function of q, as obtained at y=−1 and tem-\nperatures T/J= 0, 0.1 and 0.2. The lines are guides to the\neye.\nat finite temperatures are shown in Fig. 12.\nThe broadening at small momenta is dominated by\nthermal fluctuations and the lineshape is very similar to\nthat of the pure spin model discussed in Sec. IV. A\nfurther strong broadening in going from q= 0.3πto\nq= 0.5πsignals the relevance of coupled spin-orbital\ndegrees of freedom on the spin dynamics at intermedi-\nate and high momenta. Again, additional structures in\nS(q,ω) are visible related to the spin-orbital excitations\ndiscussed above.\nFinally, we analyze the variation of the FWHM with\nincreasing temperature for a representative value of y=\n−1, see Fig. 13. AtT >0 the thermal broadening at\nsmall momenta is clearly visible. As in the zero tempera-\nture case another strong increase of Γ qbetween q= 2π/5\nandq=π/2 is observed due to coupled spin-orbital de-\ngrees of freedom. For T/J= 0.1 andT/J= 0.2 we\nobserve that Γ qhas a temperature dependent maximum\nfrom where Γ qdecreases towards the boundary of the\nBrillouin zone. This is due to the fact that the ther-\nmal broadening of the central peaks of the dynamic spin\nstructure factor decreases from intermediate to high mo-\nmenta (see Fig. 4). Actually without coupling to any or-\nbital degrees of freedom we expect Γ qto be very small at\nthe boundary of the Brillouin zone. Thus a large band-\nwidth at q=πmakes the spin-orbital model distinct\nfrom a pure 1D Heisenberg ferromagnet.\nVI. THERMODYNAMICS\nCoupled spin-orbital degrees of freedom will not only\ninfluence the spin dynamics but also the thermodynam-\nics of the system. We expect, in particular, that the\nspin-orbital excitations which were shown to affect the/Bullet /Bullet /Bullet /Bullet\nFIG.14: Diagramatic representations ofthesecond orderco n-\ntributions to the free energy as given by Eq. (6.6).\ndynamical spin structure factor in the previous section\nwill also become observable in thermodynamic quanti-\nties when comparing the MF decoupling and the per-\nturbative solution. In order to investigate this issue we\nrewrite Eq. ( 2.1) as\nHSτ(1) =HMF+δH, (6.1)\nwithδH=HSτ(1)−HMF. The MF part reads\nHMF=N/summationdisplay\nj=1/braceleftbig\nJτ[τj·τj+1]1+JSSj·Sj+1\n−/an}bracketle{tSj·Sj+1/an}bracketri}htMF/angbracketleftbig\n[τj·τj+1]1/angbracketrightbig\nMF/bracerightbig\n.(6.2)\nThe exchange constants JS,τare defined in Eqs. ( 3.3).67\nWe use the Hamiltonian ( 6.1) to determine the free en-\nergy per site perturbatively, following the expansion,\nf=fS\nMF+fτ\nMF+1\nN/an}bracketle{tδH/an}bracketri}htc\nMF−1\n2NT/angbracketleftbig\nδH2/angbracketrightbigc\nMF+...,\n(6.3)\nwherefS\nMF(fτ\nMF) is the expression for the free energy per\nsite stemming from the spin (pseudospin) sector within\nthe MFdecouplingsolution. The subscriptindicatesthat\nthe respective correlation functions are calculated with\nHMF. Moreover the superscript cmeans that the above\nexpansion of the free energy is restricted to connected\ndiagrams. We note that Eq. ( 6.3) is a high temperature\nexpansion valid if |x|T/J≫1 and|y|T/J≫1.\nA straightforward calculation shows that the first or-\nder contribution only shifts the free energy, Eq. ( 6.3),\nand will not show up in thermodynamic observables ob-\ntained by taking derivatives of the free energy. For the\nsecond order contribution we have to evaluate two- and\nfour-point correlationfunctions both for the spin and the\npseudospin part. We use the abbreviations\n/an}bracketle{t(Sj·Sj+1)(Sl·Sl+1)/an}bracketri}ht=a+b(j,l) (6.4)\nfor the spins and\n/angbracketleftbig\n[τj·τj+1]1[τl·τl+1]1/angbracketrightbig\n=c+d(j,l) (6.5)\nfor the pseudospins. Here the site-independent quanti-\ntiesaandcstand for the disconnected parts of the four-\npoint correlation functions whereas b(j,l) andd(j,l) fol-\nlow from the connected ones. One finds after a straight-\nforward calculation that only the product of the con-\nnected parts contributes to the second order correction,14\nleading to\n/angbracketleftbig\nδH2/angbracketrightbig\nMF=J2N/summationdisplay\nj,l=1b(j,l)d(j,l).(6.6)\nTo proceed further, we again apply the MSWT to the\nspin and a Jordan-Wigner transformation to the pseu-\ndospin part. The evaluation of d(j,l) is again straight-\nforward and yields\nd(j,l) =1\n2N2/summationdisplay\nk1,k2nF[ωMF\nF(k1,0)]{1−nF[ωMF\nF(k2,0)]}\n×ei(k1−k2)(j−l){1+cos(k1+k2)}\n(6.7)\nwithωMF\nF(q,δ) as given in Eq. ( 3.6).\nThe evaluation of Eq. ( 6.4) is more involved. We first\napply a Dyson-Maleev transformation and treat the ob-\ntained expressionsusing Wick’s theorem. In addition, we\nalso have to account for the constraint of nonzero mag-\nnetization at T >0 imposed by the MSWT. The cor-\nresponding diagrams are shown in Fig. 14. Within this\napproximation the specific heat per site reads\nc=cMF+c2+... , (6.8)\nwith\nc2=J2T∂2\n∂T2N/summationtext\nj,l=1b(j,l)d(j,l)\n2NT.(6.9)\nWe calculate the first term in Eq. ( 6.8) within the MF\ndecoupling, cMF=cS\nMF+cτ\nMF, from the internal energy\nwhich is determined by the respective nearest-neighbor\ncorrelation functions allowing us to keep terms up to\nquartic order in the bosonic operators.54This strategy\nmakes it possible to obtain reliable results for cS\nMFup to\nT/(|JS|S2)≤1. The second order correction c2given in\nEq. (6.9) is obtained using the Dyson-Maleev transfor-\nmation so that quartic terms are also included and the\norder of approximation is the same. Since we are using\na high temperature expansion, Eq. ( 6.3), in combination\nwith the MSWT to evaluate the diagrams, our results\nare only valid in an intermediate temperature regime. If\nwe restrict ourselves to parameters x=−y >0 then this\ntemperature range is given by 1 /x≪T/J≪xS2. In\nthe following we therefore only consider the case x≫1\nand compare the results from perturbation theory with\nnumerical data obtained by TMRG.\nAs shown in Fig. 15, the specific heat c/(Jx)2exhibits\na broad maximum which corresponds to the character-\nistic energies of spin and fermionic particle-hole excita-\ntions. In the temperature range where the perturbative\napproach is valid we find excellent agreement with the\nnumerical solution. In particular, the perturbative cor-\nrectionc2(see Fig. 16) correctlycapturesthe weight shift\nfrom low to intermediate temperatures visible when com-\nparing the numerical and the MF decoupling solution. In0 0.2 0.4 0.60.8 1\nT/(Jx)00.20.40.6c/(4J2)00.20.40.6c/(100J2)\n0 0.2 0.4 0.60.8 1\nT/(10J)00.20.40.6c/(100J2)\n0 0.2 0.4 0.60.8 1\nT/(2J)00.20.40.6c/(4J2)(a)\n(b)cMFScMFτcMF\ncMFScMFτcMF\nFIG. 15: Specific heat per site c/(Jx)2as a function of T/Jx\nfor (a)x= 10 and (b) x= 2. The circles denote the nu-\nmerical data from TMRG with the solid line obtained by\na low-temperature fit of the TMRG data for the inner en-\nergy. The dashed lines correspond to MF decoupling while\nthe dashed-dotted lines are the results obtained by perturb a-\ntion theory. The perturbative results are expected to be val id\nfor 1/x2≪T/(Jx)≪1. Insets: Specific heat cMFwithin the\nMF solution (dashed-dotted line) with the contributions fr om\nthe spin ( cS\nMFsolid line) and the orbital ( cτ\nMFdashed line)\nsector shown separately.\nspiteofthisweightshift, theMFdecouplingyieldsoverall\na very reasonabledescription of the specific heat for both\ncases shown in Fig. 15. Forx= 10, (Fig. 15(a)) the spe-\ncific heat chas a broad maximum at T/(Jx)≃0.4. This\nmaximum results from a distinct maximum in the orbital\ncontribution cτ\nMF, see insets in Fig. 15. In contrast the\nspin contribution cS\nMFincreases steadily with increasing\ntemperature, in agreement with the higher energy scale\nfor spin excitations. As a result, the total specific heat\nhas only a weaker and broader maximum than suggested\nby the orbital part.\nThe MF decoupling does seem to fail, however, at very\nlow temperatures. Here the MF solution predicts that\nspin excitations give the dominant contribution leading15\n0 0.2 0.4 0.60.8 1\nT/(Jx)-0.0300.03c2/(Jx)2\nFIG. 16: (Color online) Perturbative contribution to the sp e-\ncific heat per site c2/(Jx)2, Eq. (6.9), as a function of T/Jx\nforx= 10 (solid line) and x= 2 (dashed line).\nto ac(T)∼√\nTbehavior. This is in contrast to an ex-\ntrapolation of the numerical data shown as dot-dashed\nlines in Fig. 15which suggests an approximately linear\ndependenceontemperature. Thisdiscrepancycomesasa\nsurprise because our perturbative calculations of the dy-\nnamical spin-structure factor in Sec. VBlead us to the\nconclusion that the magnons survive as sharp quasipar-\nticles at low energies. More generally, one might argue\nthat the ground state does not show any entanglement\nbetween the two sectors because of the classical nature\nof the FM state thus allowing for spin-wave excitations\nat low energies. From the point of view of a low-energy\neffective field theory, however, the situation is much less\nclear. While a FM chain is described by a low-energy\neffective theory with dynamical critical exponent z= 2,\nthe fermionic orbital chain has z= 1. A coupling of\nspatial spin deviations to time-dependent orbital fluc-\ntuations then seems to require that the low-energy ef-\nfective theory for the coupled system has z= 1. As a\nconsequence the temperature dependence of c(T) would\nindeed be linear. However, such an approach leaves open\nthe role and treatment of the Berry phase terms.\nWithin the perturbative approach the open question\nis whether or not higher order contributions to the self-\nenergy might induce a significant broadening of the dy-\nnamical spin-structure factor also at low energies. In this\nregard we note that the vertex responsible for the broad-\nening of S(q,ω) studied in Sec. VBdoes not play any\nrolefor the thermodynamics ofthe system. Here the con-\nstraint of vanishing magnetization means that such dia-\ngrams do not contribute to staticcorrelation functions so\nthat the lowestordercorrectionsarecaused by the vertex\nshown in Fig. 14.VII. CONCLUSIONS\nIn summary, we have investigated coupled spin-orbital\ndegrees of freedom in a one-dimensional model. For fer-\nromagnetic exchange we have shown that the considered\nmodelatspecialpointsinparameterspacecanbewritten\nin terms ofDiracexchangeoperatorsfor spin Sand pseu-\ndospinτ. As a consequence, three collective excitations\nof spin, orbital and coupled spin-orbital type do exist. In\nparticular, we discussed the case of Dirac exchange oper-\nators for S=τ= 1/2 where the dispersions of all three\nelementary excitations are degenerate and lie within the\nspin-orbital continuum. While the spin-orbital excita-\ntions stay confined in this case, a decay becomes possible\nonce we move away from this special point.\nFor antiferromagnetic exchange the one-dimensional\nspin-orbital model captures fundamental aspects of\nphysics relevant for transition metal oxides and, as we\nhave shown, sharp excitations do not exist. To address\nthe question how the spin dynamics is influenced by fluc-\ntuatingorbitalsweconsideredtheextremequantumlimit\nof orbitals interacting via an XY-type coupling. This al-\nlowed us to map the orbital sector onto free fermions\nusing the Jordan-Wigner transformation. In spin-wave\ntheory the spin sector is described by bosons so that our\nmodel corresponds to an effective boson-fermion model\nwhich applies for low temperatures. An analytic calcu-\nlation of the properties within a mean-field decoupling\napproachis then straightforward. Compared to a numer-\nical phase diagram based on the density-matrix renor-\nmalization group we find that the regime with ferromag-\nnetically polarized spins is stabilized by the decoupling\nprocedure. Furthermore, the mean-field decoupling gives\nrise to a finite temperature dimerized phase for certain\nparameterswhen starting from the ferromagnetic ground\nstate. Whileaphasewithlong-rangedimerorderatfinite\ntemperatures is not possible in a purely one-dimensional\nmodel, the mean-field approach also completely ignores\nany kind of coupled spin-orbital excitations.\nThus we developed a self-consistent perturbative\nscheme to explore the role played by spin-orbital cou-\npling. In perturbation theory the boson-fermion inter-\naction does not produce any infrared divergencies in the\none-dimensional model due to the lack of nesting. This\nmakes a perturbative calculation of the spin structure\nfactorS(q,ω) possible. At large momenta q, we find that\nS(q,ω) shows a significant broadening due to scattering\nof magnons by orbital excitations. For small momenta,\non the other hand, no broadening in this lowest order\nperturbative approach is observed because the magnon\ncannot scatter on these excitations. The onset of the\nbroadening occurs at momenta where the magnon enters\nthe boson-fermion spectrum. This point, as well as de-\ntails of the full width at half maximum is determined by\nthe strength of interaction. Most interestingly, S(q,ω)\ndoes show additional peaks and shoulders corresponding\ntocharacteristicspin-orbitalexcitations. Atpointswhere\nthe renormalized spin-wave dispersion and the dispersion16\nof these excitations cross, Kohn anomalies do occur.\nFurthermore, we compared numerical data for the spe-\ncific heat of the spin-orbital model with the mean-field\ndecoupling solution and an approach where we also took\nthe second order correction to the mean-field result into\naccount. Overall, we found that the mean-field decou-\npling does describe the specific heat reasonably well. A\nredistribution of entropic weight from low to intermedi-\nate temperatures observed when comparing the numeri-\ncal data and the mean-field solution is very well captured\nby the second order perturbative correction. An interest-\ning open point is the behavior of the specific heat c(T) at\nlow temperatures. While the mean-field solution predicts\nc(T)∼√\nTdue to spin-wave excitations, the numerical\ndata suggest instead that c(T)∼T. We have speculated\nthat a coupling of the two sectors might indeed lead to a\nlow-energy effective theory with dynamical critical expo-\nnentz= 1 but details of such a theory need to be worked\nout in the future.\nIn conclusion, we have shown that while collective\nspin-orbital excitations with infinite lifetime do not ex-\nist forantiferromagneticsuperexchangethe coupled spin-\norbital degrees of freedom have a strong influence on the\nspin excitation spectrum as well as on the thermody-\nnamic properties of the system. However, the treatment\nof spin-orbital systems beyond the range of validity of\nmean-field decoupling and perturbative schemes remains\nan open problem in theory.\nAcknowledgments\nThe authors thank O. P. Sushkov for valuable dis-\ncussions. A.M.O. acknowledges support by the Foun-\ndation for Polish Science (FNP) and by the Polish Min-\nistry of Science and Higher Education under Project No.\nN202069639. J.S. acknowledgessupport by the graduate\nschool of excellence MAINZ (MATCOR).\nAppendix: Details of the perturbative approach for\nthe boson-fermion model\nWe start bya MF decoupling, rewritingthe interaction\nas\nH1=H1,MF+(H1−H1,MF)/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright\nδH,\nwith\nH1,MF=1\nN2/summationdisplay\nkωBF(k,k,q)2/bracketleftBig\n˜nb,kf†\nqfq+ ˜nf,qb†\nkbk/bracketrightBig\n.\nWe treat δHas perturbation. The averages ˜ na,p=/angbracketleftbig\na†\npap/angbracketrightbig\nare determined self-consistently within this MF\nscheme. This leads to a renormalization of the magnon\nand fermion dispersions. We find\nζ(q) = 2JS/parenleftBigg\n|y|−1\nN/summationdisplay\nkcosk˜nf,k/parenrightBigg\n(1−cosq)−µ,\n(A.1)and\nωF(q) =/parenleftBigg\nS2+x+2\nN/summationdisplay\nk(1−cosk)˜nb,k/parenrightBigg\ncosq.(A.2)\nForT= 0 only the magnon dispersion is renormalized to\nζ(q) = 2JS(|y|+1/π)(1−cosq).\nBy virtue of Dyson’s equation we may calculate the\nbosonic Green’s function at T= 0.68Since the MF de-\ncoupling already takes the first order contributions into\naccount, the lowest order diagrams we obtain are of sec-\nond order. The self energy is thus approximated by\nthe proper self-energy obtained by summing up those\ndiagrams which may be composed by the second order\ndiagrams. From this we have Σ( q,ω)≈Σ2(q,ω)≡\nΣ+\nBF(q,ω)+Σ−\nBF(q,ω)+Σ2,1(q) where the diagrams are\ngivenby Σ+\nBF(q,ω) (Fig.6(a)), Σ−\nBF(q,ω) (Fig.6(b)), and\nΣ2,1(q) (Fig.6(c)). A straigthforward calculation reveals\nthat at zero temperature Σ 2,1(q) and Σ−\nBF(q,ω) vanish.\nFor Σ+\nBF(q,ω) we find the expression given in Eq. ( 5.5).\nFor finite temperatures we calculate the imaginary\ntime Green’s function. We find\nΣ±\nBF(q,ων,B) =−T2\nN2/summationdisplay\nk1,k2ω2\nBF(q,k1,k2)\n×/summationdisplay\na,bG(0)\nF(k2,ωa,F)G(0)\nF(q−k1+k2,ωb,F)\n×G(0)\nB(k1,±(ων,B−ωb,F+ωa,F))\n(A.3)\nfor the diagrams in Fig. 6(a,b), and\nΣ2,1(q) =−T2\nN2/summationdisplay\nk1,k2ωBF(q,q,k1)ωBF(k2,k2,k1)\n×/summationdisplay\na,b/bracketleftBig\nG(0)\nF(k1,ωa,F)/bracketrightBig2\nG(0)\nB(k2,ωb,B)(A.4)\nfor the diagram given in Fig. 6(c). Here we have used\nG(0)\nF(q,ωµ,F) = (iωµ,F−ωF(q))−1andG(0)\nB(q,ων,B) =\n(iων,B−ζ(q))−1as the fermionic and bosonic Matsub-\nara Green’s function for the noninteracting Hamiltonian,\nrespectively. ων,Fare the fermionic Matsubara frequen-\ncies. After performing the frequency sums we end up\nwith\nΣ2,1(q) =−1\nTN2/summationdisplay\nk1,k2ωBF(q,q,k1)ωBF(k2,k2,k1)\n×nB[ζ(k2)](1−nF[ωF(k1)])nF[ωF(k1)],\n(A.5)\nand the self-energies Σ±\nBF(q,ων,B) as given in Eqs. ( 5.7)\nand (5.9).17\n∗j.sirker@fkf.mpg.de\n1L. D. Landau, Phys. Z. Sowjetunion 3, 664 (1933).\n2H. Fr¨ ohlich, Advances in Physics 3, 325 (1954).\n3R. E. Peierls, Quantum Theory of Solids (Oxford Univer-\nsity Press, Oxford, 1955).\n4K. K. Das, Phys. Rev. Lett. 90, 170403 (2003).\n5A. Albus, F. Illuminati, and J. Eisert, Phys. Rev. A 68,\n023606 (2003).\n6M. A. Cazalilla and A. F. Ho, Phys. Rev. Lett. 91, 150403\n(2003).\n7L. Mathey, D.-W. Wang, W. Hofstetter, M. D. Lukin, and\nE. Demler, Phys. Rev. Lett. 93, 120404 (2004).\n8M. Lewenstein, L. Santos, M. A. Baranov, and\nH. Fehrmann, Phys. Rev. Lett. 92, 050401 (2004).\n9K. I. Kugel and D. I. Khomskii, Sov. Phys. JETP 52, 501\n(1980).\n10L. F. Feiner and A. M. Ole´ s, Phys. Rev. B 59, 3295 (1999).\n11Y. Tokura and N. Nagaosa, Science 288, 462 (2000).\n12Colossal Magnetoresistive Oxides , edited by Y. Tokura,\n(Gordon and Breach, Amsterdam, 2000).\n13N. N. Kovaleva, A. M. Ole´ s, A. M. Balbashov, A. Maljuk,\nD. N. Argyriou, G. Khaliullin, and B. Keimer, Phys. Rev.\nB81, 235130 (2010).\n14G. Khaliullin and S. Maekawa, Phys. Rev. Lett. 85, 3950\n(2000).\n15B. Keimer, D. Casa, A. Ivanov, J. W. Lynn, M. v. Zim-\nmermann, J. P. Hill, D. Gibbs, Y. Taguchi, and Y. Tokura,\nPhys. Rev. Lett. 85, 3946 (2000).\n16J. Hemberger, H.-A. K. von Nidda, V. Fritsch, J. Deisen-\nhofer, S. Lobina, T. Rudolf, P. Lunkenheimer, F. Lichten-\nberg, A. Loidl, D. Bruns, and B. B¨ uchner, Phys. Rev. Lett.\n91, 066403 (2003).\n17M. Cwik, T. Lorenz, J. Baier, R. M¨ uller, G. Andr´ e,\nF. Bour´ ee, F. Lichtenberg, A. Freimuth, R. Schmitz,\nE. M¨ uller-Hartmann, and M. Braden, Phys. Rev. B 68,\n060401 (2003).\n18Y. Ren, T. T. M. Palstra, D. I. Khomskii, A. A. Nugroho,\nA. A. Menovsky, and G. A. Sawatzky, Nature (London)\n396, 441 (1998).\n19Y. Ren, T. T. M. Palstra, D. I. Khomskii, A. A. Nugroho,\nA. A. Menovsky, and G. A. Sawatzky, Phys. Rev. B 62,\n6577 (2000).\n20M. Noguchi, A. Nakazawa, S. Oka, T. Arima, Y. Wak-\nabayashi, H. Nakao, and Y. Murakami, Phys. Rev. B 62,\nR9271 (2000).\n21G. Khaliullin, P. Horsch, and A. M. Ole´ s, Phys. Rev. Lett.\n86, 3879 (2001).\n22C. Ulrich, G. Khaliullin, J. Sirker, M. Reehuis, M. Ohl,\nS. Miyasaka, Y. Tokura, and B. Keimer, Phys. Rev. Lett.\n91, 257202 (2003).\n23P. Horsch, G. Khaliullin, and A. M. Ole´ s, Phys. Rev. Lett.\n91, 257203 (2003).\n24J. Sirker and G. Khaliullin, Phys. Rev. B 67, 100408\n(2003).\n25P. Horsch, G. Khaliullin, A. M. Ole´ s, and L. F. Feiner,\nPhys. Rev. Lett. 100, 167205 (2008).\n26J. Sirker, A. Herzog, A.M. Ole´ s, and P. Horsch, Phys. Rev.\nLett.101, 157204 (2008).\n27A. M. Ole´ s, P. Horsch, L. F. Feiner, and G. Khaliullin,\nPhys. Rev. Lett. 96, 147205 (2006).\n28C. Itoi, S. Qin, and I. Affleck, Phys. Rev. B 61, 6747(2000).\n29J. Sirker, Phys. Rev. B 69, 104428 (2004).\n30J. van den Brink, W. Stekelenburg, D. I. Khomskii, G. A.\nSawatzky, and K. I. Kugel, Phys. Rev. B 58, 10276 (1998).\n31Y. Q. Li, M. Ma, D. N. Shi, and F. C. Zhang, Phys Rev\nLett81, 3527 (1998).\n32This is reminiscent to the SU(4) model with AF exchange\nwhere all three elementary excitations are degenerate.\nHowever in the latter case the elementary excitations are\nof coupled spin-orbital type, whereas in the present case\nwe find elementary spin, orbital and combined spin-orbital\nexcitations.\n33This, we believe, is a consequence of a factor of 2 missed\nin the analysis of the elementary dispersions in this work.\n34P. A. M. Dirac, Principles of Quantum Mechanics (Oxford\nUniversity Press, Oxford, 1958).\n35E. Schr¨ odinger, Proceedings of the Royal Irish Academy\nSection A 47, 39 (1941).\n36H. A. Brown, Phys. Rev. B 31, 3118 (1985).\n37L.F. Feiner, A.M.Ole´ s, andJ. Zaanen, J. Phys.: Condens.\nMatter10, L555 (1998).\n38A. M. Ole´ s, L. F. Feiner, and J. Zaanen, Phys. Rev. B 61,\n6257 (2000).\n39M. Takahashi, Phys. Rev. B 42, 766 (1990).\n40M. Takahashi, Phys. Rev. B 47, 8336 (1993).\n41M. Takahashi, Phys. Rev. Lett 58, 168 (1987).\n42M. Takahashi, Prog. Theor. Phys. Suppl. 87, 233 (1986).\n43T. Mizokawa and A. Fujimori, Phys. Rev. B 54, 5368\n(1996).\n44H. Sawada and K. Terakura, Phys. Rev. B 58, 6831 (1998).\n45T. Mizokawa, D. I. Khomskii, and G. A. Sawatzky, Phys.\nRev. B60, 7309 (1999).\n46J.-H. Park, L. H. Tjeng, A. Tanaka, J. W. Allen, C. T.\nChen, P. Metcalf, J. M. Honig, F. M. F. de Groot, and\nG. A. Sawatzky, Phys. Rev. B 61, 11506 (2000).\n47R. J. Bursill, T. Xiang, and G. A. Gehring, J. Phys. Cond.\nMat.8, L583 (1996).\n48X. Wang and T. Xiang, Phys Rev B 56, 5061 (1997).\n49J. Sirker and A. Kl¨ umper, Europhys. Lett. 60, 262 (2002).\n50S. R. White and D. S. Huse, Phys. Rev. B 48, 3844 (1993).\n51S. Miyashita and N. Kawakami, J. Phys. Soc. Japan 74,\n758 (2005).\n52C. Itoi, S. Qin, and I. Affleck, Phys. Rev. B 61, 6747\n(2000).\n53Y. Chen, Z. D. Wang, Y. Q. Li, and F. C. Zhang, Phys.\nRev. B75, 195113 (2007).\n54A. Herzog, P. Horsch, A. M. Ole´ s, and J. Sirker, Journal\nof Physics Conference Series 200, 022017 (2010).\n55P. Pincus, Solid State Communications 9, 1971 (1971).\n56The discussion of the dynamical spin structure factor for\nthe dimerized FM chain will be given elsewhere.\n57We note that the MF decoupling applied to the interaction\ngiven in Eq. ( 2.8) does not renormalize the dispersion of\nthe fermions at T= 0.\n58T. Giamarchi, Quantum Physics in One Dimension\n(Clarendon Press, Oxford, 2003).\n59W. Metzner, C. Castellani, and C. D. Castro, Advances in\nPhysics47, 317 (1998).\n60W. Kohn, Phys. Rev. Lett. 2, 393 (1959).\n61E. J. Woll and S. J. Nettel, Phys. Rev. 123, 769 (1961).\n62G. Barnea and G. Horwitz, J. Phys. C 8, 2124 (1975).18\n63H. B. Møller and J. C. G. Houmann, Phys. Rev. Lett. 16,\n737 (1966).\n64S. V. Halilov, H. Eschrig, A. Y. Perlov, and P. M. Oppe-\nneer, Phys. Rev. B 58, 293 (1998).\n65M. Pajda, J. Kudrnovsk´ y, I. Turek, V. Drchal, and\nP. Bruno, Phys. Rev. B 64, 174402 (2001).\n66S. Mor´ an, C. Ederer, and M. F¨ ahnle, Phys. Rev. B 67,\n012407 (2003).\n67Performing the MF decoupling with the Hamiltonian ( 2.1)also takes the quartic orders of the Dyson-Maleev transfor-\nmation into account. These higher orders are neglected in\nthe MF decoupling performed in Sec. V. For low temper-\natures, i.e. the temperature region we have used in Sec. V\nthe contributions from these higher order terms are small.\n68A. L. Fetter and J. D. Walecka, Quantum Theory of Many-\nParticle Sytems (McGraw-Hill, Inc., New York, 1971)." }, { "title": "1912.07728v1.Spin_current_manipulation_of_photoinduced_magnetization_dynamics_in_heavy_metal___ferromagnet_double_layer_based_nanostructures.pdf", "content": "IEEE TRANSACTIONS ON MAGNETICS, VOL. 53, NO. 11, NOVEMBER 2017 1\nSpin-current manipulation of photoinduced magnetization dynamics\nin heavy metal / ferromagnet double layer based nanostructures\nSteffen Wittrock1, Dennis Meyer2, Markus M ¨uller2, Henning Ulrichs2, Jakob Walowski3, Maria Mansurova3,\nUlrike Martens3, and Markus M ¨unzenberg3\n1Unit´e Mixte de Physique CNRS/Thales, Universit ´e Paris Sud, 1 Avenue Augustin Fresnel, 91767 Palaiseau, France\n2I. Physikalisches Institut, Georg-August-Universit ¨at G ¨ottingen, Friedrich-Hund-Platz 1, 37077 G ¨ottingen, Germany\n3Institut f ¨ur Physik, Ernst-Moritz-Arndt-Universit ¨at Greifswald, Felix-Hausdorff-Str. 6, 17489 Greifswald, Germany\nSpin currents offer a way to control static and dynamic magnetic properties, and therefore they are crucial for next-generation\nMRAM devices or spin-torque oscillators. Manipulating the dynamics is especially interesting within the context of photo-magnonics.\nIn typical 3dtransition metal ferromagnets like CoFeB, the lifetime of light-induced magnetization dynamics is restricted to about 1\nns, which e.g. strongly limits the opportunities to exploit the wave nature in a magnonic crystal filtering device. Here, we investigate\nthe potential of spin-currents to increase the spin wave lifetime in a functional bilayer system, consisting of a heavy metal (8 nm\nof\f-Tantalum (Platinum)) and 5 nm CoFeB. Due to the spin Hall effect, the heavy metal layer generates a transverse spin current\nwhen a lateral charge current passes through the strip. Using time-resolved all-optical pump-probe spectroscopy, we investigate how\nthis spin current affects the magnetization dynamics in the adjacent CoFeB layer. We observed a linear spin current manipulation\nof the effective Gilbert damping parameter for the Kittel mode from which we were able to determine the system’s spin Hall angles.\nFurthermore, we measured a strong influence of the spin current on a high-frequency mode. We interpret this mode an an exchange\ndominated higher order spin-wave resonance. Thus we infer a strong dependence of the exchange constant on the spin current.\nIndex Terms —Spin Hall effect, spin current, magnetization dynamics, magnetooptical Kerr-effect.\nI. I NTRODUCTION\nTHE spin-transfer effect describes the transfer of spin an-\ngular momentum to a ferromagnet’s magnetization from\nan injected spin polarized current. Since its prediction by L.\nBerger [1], this effect has experienced a high research interest\nas it permits to manipulate and control the magnetization of a\nthin ferromagnetic (FM) layer. Especially when the resulting\nspin transfer torque is collinear with the damping torque,\nmagnetic dissipation can be controlled. Thereby the life time\nof spin wave dynamics can be drastically enhanced.\nAmong the possible methods of creating the necessary spin\ncurrent, the exploitation of the spin Hall effect has become a\npowerful mean since its first observation only a decade ago\n[2], [3], [4], [5]. Governed by spin-orbit coupling phenomena,\nit generates a spin current jsfrom a transverse charge current\njewithout any need for neither a ferromagnet nor an external\nmagnetic field. The efficiency of the conversion process can\nbe described by the spin Hall angle (SHA) \u0002SH=js=je.\nHere, we investigate the photo-induced magnetization dy-\nnamics in a few nanometer thin soft magnetic layer consisting\nof amorphous metallic cobalt iron boron alloy (Co 20Fe60B20)\nunder influence of a strong spin current generated by the SHE\nin an adjacent heavy metal film made of platinum (Pt) or\ntantalum (Ta). The sputter conditions for Ta were chosen such\nthat the film has grown in the high resistive \f-phase for which\na high SHA has recently been reported [6].\nII. E XPERIMENTAL\nThe samples consist of two functional thin layers (fig. 1a):\n8nm Pt or\f-Ta as a SHE-material generating the spin current\nManuscript received April 21, 2017; revised April 21, 2017. Corresponding\nauthor: S. Wittrock (email: steffen.wittrock@u-psud.fr).and5nm of ferromagnetic amorphous CoFeB, into which\nthe spin current is being injected in order to manipulate its\nmagnetization dynamics. The layer stack is complemented by\n3nm of Ru as a capping layer. CoFeB and Ta are deposited by\nargon ion sputtering and Pt and Ru by e-beam evaporation. All\npreparation steps are conducted in situ in ultra-high vacuum.\nSubsequent structuring of the samples by e-beam lithography\nenables the electrical contacting and generation of a high\ncharge current density (fig. 1b).\n(a)\n(b)\n (c)\nFig. 1: Experimental characteristics. (a) Schematic processes\nin the functional layer stack of a SHE material (blue) and the\nferromagnetic CoFeB (yellow). (b) Patterned sample structure,\nthe widthxwas12\u0016m for the results shown here. The red\nmarked area was excited by the laser spot. (c) Pump-probe\nsetup with ratio of powers Ppump :Pprobe = 95 : 5 , time\nresolution is realized by a delay stage, a double modulation\ntechnique of photoelastic modulator (PEM) and chopper fre-\nquency was used [7].\nWe used time resolved pump-probe spectroscopy exploitingIEEE TRANSACTIONS ON MAGNETICS, VOL. 53, NO. 11, NOVEMBER 2017 2\nthe magnetooptical Kerr effect to excite and measure magne-\ntization dynamics (schematical setup in 1c). The laser pulses\nwith central wave length \u0015= 800 nm have an autocorrelation\nlength of \u0001\u001c\u001980fs and a repetition rate of 250kHz. The\npump spot size was \u001960\u0016m providing an optical fluence of\nF= 15 mJ=cm2.\nThe incoming pump pulse induces the dynamics at \u001c= 0\nby firstly generating hot electrons, which thermalize on a\ntimescale of \u001c\u0018100fs due to electron-electron-scattering.\nFurther scattering events with phonons and spins lead to\nenergy transfer into the phonon and spin system giving rise to\nultrafast demagnetization [8]. Caused by the high change in\ntemperature on the time scale of \u00181ps, the local anisotropy\nconstant of the ferromagnetic material and thus the effective\nmagnetic field ~Heffis changed. With ~Heffrereaching its\nequilibrium position after a few picoseconds, the magnetiza-\ntion dynamics corresponding to the Landau-Lifshitz-Gilbert-\nequation (LLG) is excited. In order to subsequently trigger\na coherent precession of the magnetization, an external field\n(145mT) was applied at an out-of-plane angle of '= 35\u000e,\ntransversal to ~je.\nIII. R ESULTS -NANOSECOND TIMESCALE\n-0.5-0.4-0.3-0.2-0.10.00.10\n2 004 006 008 001 000-0.04-0.020.000.020.040.06T\nime delay τ [ps]TRMOKE signal [a.u.] \nData \nBackroundµ0H = 145 mT ; ϕ = 35° ; j = 6.2e10 A/m2 \nbackround-cleaned, fitted data \nDamping ∼ exp(- τ/τα)\n(a)\n-10-8-6-4-2024681012.012.112.212.312.412.512.612.712.812.9Kittel-Frequency [GHz]C\nurrent density in Ta-layer [1010 A/m2]µ0H = 145 mT ; ϕ = 35° \n(b)\n-10-8-6-4-202468101.301.351.401.451.501.551.601.65µ0MS [T]C\nurrent density in Ta-layer [1010 A/m2]µ0H = 145 mT ; ϕ = 35° (c)\nFig. 2: (a) Typical measurement data and Kittel mode after\nsubstraction of exponential background. The TRMOKE-signal\nis proportional to the magnetization. Dependence of (b) Kittel\nfrequency and (c) saturation magnetization on the current\ndensity in the Ta-layer.\nBesides the dominating coherent, in spatially homogeneous\ngeometries called Kittel mode oscillations, also incoherent\nphonons and magnons contribute to the signal and give rise\nto a certain background [9], [8], which can be modelled by\nexponential functions. Substracting it from the raw data (fig.\n2) highlights the magnetic oscillations. These are analysedwith respect to frequency and damping for different current\ndensities passing through the SHE-material.\nA. Kittel frequencies, magnetization, and Oersted field\nThe dependence of the Kittel frequency on jis shown in\nfig. 2b. The mainly parabolic behaviour can be attributed to\nthe reduction of the saturation magnetization due to Joule-\nheating. The observable asymmetries for opposite current\ndirections can be related to the Oersted field produced by the\ncurrent, or to the presence of a field-like torque. Analysing\nthe asymmetries, an Oersted field of HOe=ajwitha=\n(1:23\u00060:04)\u000110\u00008m is determined which agrees well with\ntheoretical estimations. Even if present, the field-like torque\nmust be much smaller than the effect arising from the Oersted\nfield. Taking also into account the in-plane applied magnetic\nfield component Hx, the saturation magnetization can be deter-\nmined from the Kittel formula !=\r\u00160p\nHx(Hx+MS)(fig.\n2c). Note that for the given field geometry, the magnetization’s\nout-of-plane component is only around 2%, as is estimated\nby micromagnetic simulations; therefore this approximation\nis valid. Besides Joule-heating, also the energy deposition of\nthe pump pulse heats up the sample locally to around 400-\n450K at a time scale of up to 1ns. Thus the saturation\nmagnetization at j= 0 A/m2is slightly lower than the\nroom temperature value of \u00160MS= 1:63T, determined by a\nvibrating sample magnetometer [10]. The Joule-heating effect\nleads to a temperature increase especially in Ta of up to 200K\nat a maximum current density of jTa= 9:3\u00011010A/m2.\nB. Effective Gilbert damping parameter of the Kittel mode\nFrom the exponential decay time \u001c\u000bof the Kittel mode (fig.\n2a) and taking into account the full in-plane magnetic field\nHx=Hextcos(') +HOe, and the current dependence of\nthe magnetization, we calculate the effective Gilbert damping\nparameter using:\n\u000bKittel =\u0012\n\u001c\u000b\r\u00160\u0012\nHx+MS\n2\u0013\u0013\u00001\n: (1)\n-10-8-6-4-202468100.0050.0060.0070.0080.0090.0100.011Effective Gilbert dampingC\nurrent density in Ta-layer [1010 A/m2] Calculated data \nLinear fit\n(a)\n-30- 20- 100 1 02 00.000.010.020.030.04C\nurrent density in Pt-layer [1010 A/m2]Effective Gilbert damping Calculated data \nLinear fit (b)\nFig. 3: Current dependence of the effective Gilbert damping\nparameter for (a) the Ta based sample and (b) the Pt based\nsample. From the linear fit, the SHE efficiency can be ex-\ntracted.\nResults for the two SHE-materials Ta and Pt are shown in\nfig. 3. We determined the SHA \u0002SHby fitting the formula:\n~\u000b=\u000b+j\u0001\u0002SH\u0001~\n2ed\u00160MSHx: (2)IEEE TRANSACTIONS ON MAGNETICS, VOL. 53, NO. 11, NOVEMBER 2017 3\nNote that this expression was derived from linearizing the\nLandau-Lifshitz-Gilbert-Slonczewski-equation (LLGS).\nAveraging a few datasets gives the SHAs of:\n\u0002Ta\nSH=\u00000:043\u00060:011;\nand\n\u0002Pt\nSH= 0:086\u00060:012:\nThese results lie within the values obtained by other groups\n[11] and in particular confirm the different sign to be expected\nfor Pt [12], [13] and Ta [14], [6].\nC. Spin-pumping effect\nThe spin mixing conductance g\"#of a layer combination\ncan be estimated through comparison of the effective damping\nparameters ~\u000bof two layer stacks involving the same ferromag-\nnetic material. Using Ta and Pt as SHE-materials and CoFeB\nas the ferromagnet, the spin mixing conductance of Ta can be\ncalculated using:\ng\"#\nTa=\u00004\u0019MSd\u0001\u0001~\u000b\n~\r+g\"#\nPt (3)\n\u0001~\u000b= ~\u000bPt\u0000~\u000bTadenotes the difference of the effective\nmagnetic damping constants of the d= 5 nm thick CoFeB\nlayer for the two different adjacent SHE-materials. Average\nvalues of a few measurements give ~\u000bPt= (1:20\u00060:15)\u000110\u00002\nand~\u000bTa= (0:69\u00060:05)\u000110\u00002forj= 0. The Pt/CoFeB spin\nmixing conductance of g\"#\nPt= (4\u00061)\u00011019m\u00002[18] was used\nas a reference value.\nWe evaluate the spin mixing conductance of Ta/CoFeB to\ng\"#\nTa= (1:9\u00061:2)\u00011019m\u00002. The smaller value for Ta/CoFeB\ncompared to Pt/CoFeB is expected as similar values were\nalready obtained in Ta/NiFe [15] resp. Pt/NiFe [16] bilayers.\nIn conclusion, this analysis shows that TRMOKE is also a\nsuitable method to determine the spin mixing conductance\nwhich is an important parameter for heavy metal/FM based\nbilayer systems.\nIV. R ESULTS -PICOSECOND TIMESCALE\nOn the timescale of a few picoseconds, when the magneti-\nzation starts relaxating again, we observed a strongly damped,\nultrafast oscillation in the terahertz regime (fig. 4c), usually\nexistent for one or two periods (fig. 4). We identified this os-\ncillation as the well-known perpendicular standing spin-wave\n(PSSW) mode of first order n= 1. Analysing its frequencies\n(fig. 4c) and taking into account the saturation magnetization\ngathered on the nanosecond timescale, especially the exchange\nstiffnessA(fig. 4d) can be obtained from:\n!=\r\u00160s\u0012\nHx+2A\n\u00160MSk2\u0013\u0012\nHx+2A\n\u00160MSk2+MS\u0013\n(4)\nwithk2=k2\nz= (n\u0019=d)2,n2Nthe quantized wavevector\nnormal to the plane. The exchange stiffness is shown in fig.\n4d and found to depend on the spin current.Note that pinning at interfaces can alter the effective wave\nlength of the exchange mode. Without further experimental\nevidence, assuming zero pinning is the simplest model. Since\nthe exchange constant fits to known values determined from\nTRMOKE experiments on thicker films [17], we stick to this\nmodel.\n-2-10123456789-0.5-0.4-0.3-0.2-0.10.00.1 \nsmoothed data \n data pointsTRMOKE signal [a.u.]T\nime delay τ [ps]Demagnetizationµ0H = 145 mT ; ϕ = 35° ; j = 6.2e10 A/m2\n(a)\n34 5 6 7 8 9 -0.06-0.04-0.020.000.020.040.06 original data \nsmoothed data \ndamped sinus fitTRMOKE signal [a.u.]T\nime delay τ [ps]\n(b)\n-10-8-6-4-20246810520540560580600620640Frequency [GHz]C\nurrent density in Ta-layer [1010 A/m2] (c)\n-8-6-4-2024682.32.42.52.62.72.82.93.03.13.23.3Exchange stiffness A [10-11 J/m]C\nurrent density in Ta-layer [1010 A/m2]\n(d)\n-8-6-4-202 4 6 8 0.020.040.060.080.100.120.140.16C\nurrent density in Ta-layer [1010 A/m2]αPSSW (e)\nFig. 4: (a) The first ten picoseconds of the excited dynamics\nincluding an ultrafast, highly damped oscillation during re-\nmagnetization (b). (c-e) Current dependence of certain derived\nparameters.\nThe Gilbert damping parameter of the PSSW mode (fig.\n4e) can again be determined from the oscillation’s exponential\ndecay time\u001c\u000b:\n\u000bPSSW =\u0012\n\u001c\u000b\r\u00160\u0012\nHx+2Ak2\n\u00160MS+MS\n2\u0013\u0013\u00001\n:(5)\nEspecially the exchange term \u0018Ak2=MSproves to be dom-\ninant due to the small CoFeB thickness. We attribute the (at\nleast for positive jTa) linear behaviour of \u000bPSSW to the SHE.\nThe curve’s slope (fig. 4e, jTa\u00150) possesses the same sign\nas the Kittel mode damping \u000bKittel . Furthermore, \u000bPSSW\nis found to be one order of magnitude higher than \u000bKittel .\nThe relative change from \u000bPSSW\u00190:04for a high positive\ncurrent density up to \u000bPSSW\u00190:13forj= 0shows a strong\ndependence on the spin current which is around three times\nhigher than for the Kittel mode.\nV. D ISCUSSION\nAccording to our experiments, the SHA for a Pt-based layer\nstructure is almost double the SHA of the Ta-system. A strongIEEE TRANSACTIONS ON MAGNETICS, VOL. 53, NO. 11, NOVEMBER 2017 4\nJoule-heating effect occurred especially in the Ta based struc-\nture because of its high resistivity. The latter also contributes\nto an eventual systematic reduction of the tantalum’s SHA by\nup to 10% assuming the capped Ru displays a small negative\nSHA of \u0002SH\u0019\u00000:001[19]. This would lead to a spin current\nwith opposing sign flowing into the CoFeB from above.\nThe strong spin pumping mechanism and thus increase of\nthe magnetic damping constant especially in the Pt sample\nis theoretically expected due to platinum’s high spin flip\nprobability. Adding an interlayer with high spin diffusion\nlength (such as Cu [20], [21], [22]) could solve this issue.\nThe interesting discovery of the fast, highly damped mag-\nnetic oscillation within the first 10ps of the remagnetization\nprocess is identified as the PSSW mode of first order. It\nshows a strong dependence on the injected spin current which\ninfluences its frequency, the exchange stiffness and the mode’s\ndamping. Note that usually, the PSSW mode is optically\nexcitable and observable in samples, whose layer thickness\nis higher than the optical penetration depth which is \u001830nm\nin our case. To obtain clear evidence of this mode’s properties,\nfurther investigations with enhanced signal to noise ratio are\nnecessary.\nVI. C ONCLUSION AND OUTLOOK\nIn this article we discussed the photoinduced magnetization\ndynamics under influence of an injected spin current, generated\nby the SHE. The powerful tool of time resolved magnetoop-\ntical Kerr effect, compared to more established methods like\nBLS or ST-FMR, allowed to investigate nanosecond as well\nas (sub-)picosecond dynamics and to get insights into high-\nfrequency and non-equilibrium dynamics so far not in the\nfocus within this context.\nWe discussed the influence of Joule-heating on our samples\nand described the linear manipulation of the Kittel mode’s\nGilbert damping through a spin current. The spin Hall angles\nwhich could be determined for the two different, commonly\nused SHE materials Ta and Pt, lie within other reported values\n[11].\nWe reported a magnetic oscillation at the picosecond\ntimescale which is found to be highly dependent on the spin\ncurrent. The exact behaviour of this mode has to be further\ninvestigated in future experiments with enhanced signal to\nnoise ratio. Furthermore, the interesting timescales of the de-\nmagnetization process will be adressed in future publications.\nStandard methods such as ST-FMR do not allow to address\nsuch high-frequency dynamics easily. Our approach enables\nus to enter this temporal regime, which provides new insights\non the action of spin torques on picosecond time scales.\nACKNOWLEDGMENT\nThe authors acknowledge financial support by the DFG,\nwithin the CRC 1073 ’Atomic scale control of energy con-\nversion’.\nREFERENCES\n[1] B ERGER , L.: Exchange interaction between ferromagnetic domain wall\nand electric current in very thin metallic films. Journal of Applied Physics\n55, 6, p. 1954-1956, 1984.[2] K ATO, Y.K.; M YERS , R.C.; G OSSARD , A.C.; A WSCHALOM , D.D.:\nObservation of the Spin Hall Effect in Semiconductors. Science 306, 5703,\np. 1910-1913, 2004.\n[3] W UNDERLICH , J.; K ¨ASTNER , B.; S INOVA , J.; J UNGWIRTH , T.: Exper-\nimental discovery of the spin-Hall effect in Rashba spin-orbit coupled\nsemiconductor systems. condmat/0410295.\n[4] D AY, C.: Two Groups Observe the Spin Hall Effect in Semiconductors.\nPhys. Today 58, 2, 17, 2005.\n[5] W UNDERLICH , J.; K ¨ASTNER , B.; S INOVA , J.; J UNGWIRTH , T.: Experi-\nmental observation of the spin-Hall effect in a two-dimensional spin-orbit\ncoupled semiconductor system. Physical Review Letters, 94(4), 047204,\n2005.\n[6] L IU, L.; P AI, C.-F.; L I, Y.; T SENG , H. W.; R ALPH , D. C.; B UHRMAN ,\nR. A.: Spin-torque switching with the giant spin Hall effect of tantalum.\nScience 336, 6081, 2012.\n[7] K OOPMANS , B.: Laser-Induced Magnetization Dy-\nnamics. In: H ILLEBRANDS , B.; O UNADJELA , K.:\nSpin Dynamics in Confined Magnetic Structures II. , 2003.\n[8] B EAUREPAIRE , E.; M ERLE , J.-C.; D AUNOIS , A.; B IGOT , J.-Y.: Ultra-\nfast Spin Dynamics in Ferromagnetic Nickel. Phys. Rev. Lett. 76, p. 4250-\n4253, 1996.\n[9] D JORDJEVIC , M.; L ¨UTTICH , M.; M OSCHKAU , P.; G UDERIAN , P.;\nKAMPFRATH , T.; U LBRICH , R. G.; M ¨UNZENBERG , M.; F ELSCH , W.;\nMOODERA , J. S.: Comprehensive view on ultrafast dynamics of ferro-\nmagnetic films. Physica status solidi (c) 3, 5, p. 1347-1358, 2006.\n[10] M ANSUROVA , M.: Magnetization and elastic dynamics in nanostruc-\ntured metamaterials. I. Physikalisches Institut, Georg-August-Universit ¨at\nG¨ottingen, 2016.\n[11] S INOVA , J.; V ALENZUELA , S.O.; W UNDERLICH , J.; B ACK, C.H.;\nJUNGWIRTH , T.: Spin Hall effects. Reviews of Modern Physics 87, 4,\np. 1213-1260, 2015.\n[12] L IU, L.; M ORIYAMA , T.; R ALPH , D.C.; B UHRMAN , R.A.: Spin-Torque\nFerromagnetic Resonance Induced by the Spin Hall Effect. Phys. Rev.\nLett. 106, 036601, 2011.\n[13] A NDO , K.; T AKAHASHI , S.; H ARII, K.; S ASAGE , K.; I EDA, J.;\nMAEKAWA , S.; S AITOH , E.: Electric Manipulation of Spin Relaxation\nUsing the Spin Hall Effect. Phys. Rev. Lett. 101, 036601, 2008.\n[14] M OROTA , M.; N IIMI, Y.; O HNISHI , K.; W EI, D.H.; T ANAKA , T.;\nKONTANI , H.; K IMURA , T.; O TANI , Y.: Indication of intrinsic spin Hall\neffect in 4 dand 5 dtransition metals. Phys. Rev. B 83, 174405, 2011.\n[15] M ONTOYA , E.; O MELCHENKO , P.; C OUTTS , C.; L EE-HONE, N.R.;\nH¨UBNER , R.; B ROUN , D.; H EINRICH , B.; G IRT, E.: Spin transport in\ntantalum studied using magnetic single and double layers. Phys. Rev. B\n94, 054416, 2016.\n[16] C ZESCHKA , F.D.; D REHER , L.; B RANDT , M.S.; W EILER , M.; A L-\nTHAMMER , M.; I MORT , I.-M.; R EISS, G.; T HOMAS , A.; S CHOCH , W.;\nLIMMER , W.; H UEBL , H.; G ROSS , R.; G OENNENWEIN , S.T.B.: Scaling\nbehavior of the spin pumping effect in ferromagnet/platinum bilayers.\narXiv:1012.3017 , 2011.\n[17] U LRICHS , H; L ENK, B.; M ¨UNZENBERG , M.: Magnonic spin-wave\nmodes in CoFeB antidot lattices. Appl. Phys. Lett. 97, 092506, 2010.\n[18] R UIZ-CALAFORRA , A.; B R¨ACHER , T.; L AUER , V.; P IRRO , P.; H EINZ ,\nB.; G EILEN , M.; C HUMAK , A.V.; C ONCA , A.; L EVEN , B.; H ILLE -\nBRANDS , B.: The role of the non-magnetic material in spin pumping and\nmagnetization dynamics in NiFe and CoFeB multilayer systems. Journal\nof Applied Physics 117, 16, 2015.\n[19] K AMPFRATH , T.; B ATTIATO , M.; M ALDONADO , P.; E ILERS , G.;\nN¨OTZOLD , J.; M ¨AHRLEIN , S.; Z BARSKY , V.; F REIMUTH , F.;\nMOKROUSOV , Y. ; B L¨UGEL , S.; W OLF, M.; R ADU , I.; O PPENEER ,\nP.M. ; M ¨UNZENBERG , M.: Terahertz spin current pulses controlled by\nmagnetic heterostructures. Nature nanotechnology 8, 4, p. 256260, 2013.\n[20] M IZUKAMI , S.; A NDO , Y.; M IYAZAKI , T.: Effect of spin diffusion on\nGilbert damping for a very thin permalloy layer in Cu/permalloy/Cu/Pt\nfilms. Phys. Rev. B 66, 104413, 2002.\n[21] S UN, Y.; C HANG , H.; K ABATEK , M.; S ONG , Y.-Y.; W ANG , Z.;\nJANTZ , M.; S CHNEIDER , W.; W U, M.; M ONTOYA , E.; K ARDASZ , B.;\nHEINRICH , B.; V ELTHUIS , S.G.E.; S CHULTHEISS , H.; H OFFMANN , A.:\nDamping in Yttrium Iron Garnet Nanoscale Films Capped by Platinum.\nPhys. Rev. Lett. 111, 106601, 2013.\n[22] R OJAS -S´ANCHEZ , J.-C.; R EYREN , N.; L ACZKOWSKI , P.; S AVERO ,\nW.; A TTAN ´E, J.-P.; D ERANLOT , C.; J AMET , M.; G EORGE , J.-M.; V ILA,\nL.; J AFFR `ES, H.: Spin Pumping and Inverse Spin Hall Effect in Platinum:\nThe Essential Role of Spin-Memory Loss at Metallic Interfaces. Phys. Rev.\nLett. 112, 106602, 2014." }, { "title": "1608.08153v1.Spin_pumping_and_measurement_of_spin_currents_in_optical_superlattices.pdf", "content": "Spin pumping and measurement of spin currents in optical superlattices\nC. Schweizer1;2, M. Lohse1;2, R. Citro3;4, I. Bloch1;2\n1Fakultät für Physik, Ludwig-Maximilians-Universität, Schellingstrasse 4, D-80799 München, Germany\n2Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Strasse 1, D-85748 Garching, Germany\n3Dipartimento di Fisica “E. R. Caianiello”, Università degli Studi di Salerno,\nVia Giovanni Paolo II 132, I-84084 Fisciano (Salerno), Italy\n4SPIN-CNR Salerno, Via Giovanni Paolo II 132, I-84084 Fisciano (Salerno), Italy\n(Dated: September 11, 2021)\nWe report on the experimental implementation of a spin pump with ultracold bosonic atoms in\nan optical superlattice. In the limit of isolated double wells it represents a 1D dynamical version\nof the quantum spin Hall effect. Starting from an antiferromagnetically ordered spin chain, we\nperiodically vary the underlying spin-dependent Hamiltonian and observe a spin current without\ncharge transport. We demonstrate a novel detection method to measure spin currents in optical\nlattices via superexchange oscillations emerging after a projection onto static double wells. Further-\nmore, we directly verify spin transport through in-situ measurements of the spins’ center of mass\ndisplacement.\nExposing materials to strong magnetic fields has led to\nremarkable discoveries, most prominently the pioneering\nobservation of the integer and fractional quantum Hall\neffect [1, 2]. These quantum phenomena surprise due\nto their robustness and independence of material prop-\nerties, arising from their topological nature [3]. In this\ncontext, Thouless recognized that 1D dynamical systems\ncan share the same topological character as the 2D in-\nteger quantum Hall (IQH) effect [4, 5]. Such topological\nchargepumpsexhibitaquantizedtransportperpumpcy-\ncle in a gapped filled band of an adiabatically and period-\nically evolving potential. More recently, a fundamentally\ndifferent quantum state was observed [6], the topologi-\ncal insulator (TI) [7, 8], which preserves in addition to\ncharge time-reversal symmetry. In 2D systems with spin\nconservation, it exhibits the quantum spin Hall (QSH)\neffect, characterized by a quantized spin but vanishing\ncharge conductance. Analogous to the Thouless pump, a\ndynamical version of a TI can be designed – a quantum\nspin pump [9–11].\nSpin pumps could serve as spin current sources, e.g. for\nspintronic applications [12]. Such spin current generators\nhave been proposed based on the spin Hall effect [13, 14],\nthe cyclic variation of two system parameters in inter-\nacting quantum wires [15, 16], and topological insula-\ntors [17]. However, spin pumps have been realized only\nin few experiments, e.g. in quantum dot structures [18]\nand by parametrically excited exchange magnons [19].\nHere, we demonstrate the first implementation of a spin\npump with ultracold bosonic atoms in optical superlat-\ntices and present a direct measurement of the arising spin\ncurrent.\nIn1983,Thoulessinvestigatedparticletransportintwo\nsuperimposed1Dperiodicpotentialsadiabaticallymoved\nrelative to each other . This sliding motion periodically\nvaries the combined potential and thereby the underlying\nsingle-particle Hamiltonian, which can be parametrized\nby the cyclic pump parameter \u001e. During the pumping, aparticle acquires an anomalous velocity proportional to\ntheBerrycurvaturedefinedontheclosedsurfacespanned\nby\u001eand the quasi-momentum k. The resulting displace-\nmentafteronepumpcycleonlydependsonthegeometric\nproperties of the pump cycle and is not quantized unless\nall quasi-momenta of a band are occupied equally. In\nthis case, the pump is topological with a displacement\nproportional to the Chern number, the integral of the\nBerry curvature over the entire surface. Thus, transport\nis quantized and robust against perturbations [4, 5]. Re-\ncently, such geometric and quantized, topological pumps\nhave been realized with ultracold bosonic [20, 21] and\nfermionic atoms [22].\nIn analogy to the Thouless pump, a spin pump can\nbe thought of as a dynamical version of a QSH system\n[9], which is characterized by a bulk excitation gap and\ngapless edge excitations. In general, the electron spin\nis not conserved, e.g. in the presence of spin-orbit cou-\npling, and therefore unconventional topological invari-\nants, like the Z2index [7], are needed for classification.\nNon-interacting QSH systems with spin conservation can\nbe interpreted as two independent IQH systems. There-\nfore, a quantum spin pump can be composed of two in-\ndependent pumps, where the up and down spins have\ninverted Berry curvature and are therefore transported\nin opposite directions.\nA quantum spin pump can be implemented with ultra-\ncold atoms in two hyperfine states in a spin-dependent\ndynamically controlled optical superlattice, which can be\nformed by superimposing two lattices with periods ds\nanddl= 2ds. In the tight binding limit, a spin in this\nsuperlattice is described by the Rice-Mele model [23],\nwhich comprises staggered on-site energies \u0006\u0001=2be-\ntween neighboring sites and alternating tunnel cou-\nplings1\n2(J\u0006\u000eJ)with the dimerization parameter \u000eJ.\nPumping can be induced by an adiabatic modulation\nof the potential and corresponds to a loop in parame-\nter space (\u000eJ,\u0001) around the degeneracy point ( \u000eJ= 0,arXiv:1608.08153v1 [cond-mat.quant-gas] 29 Aug 20162\n\u0001 = 0). If the two spin components do not interact\nwith each other, their pumping motion is independent\nand a spin pump can be realized by a spin-dependent\ndeformation of the potential, so that time-reversal sym-\nmetry is retained and their Berry curvature is reversed.\nThe associated spin transport is quantized only for equal\noccupation of all quasi-momenta, which can be realized\nwith non-interacting fermions by placing the Fermi en-\nergy in the band gap and bosons by localizing each spin\ncomponent to a Mott insulator with negligible inter-spin\ninteraction.\nIn addition, the two spin components can be coupled\nby introducing on-site interactions Ubetween the atoms.\nFor hardcore interactions and unit filling, the bare tun-\nneling is suppressed and the system can be described by\na 1D spin chain\n^H=\u00001\n4X\nm\u0000\nJex+ (\u00001)m\u000eJex\u0001\u0010\n^S+\nm^S\u0000\nm+1+h.c.\u0011\n+\u0001\n2X\nm(\u00001)m^Sz\nm(1)\nwithspin-dependenttilt \u0001andalternatingexchangecou-\npling1\n2(Jex\u0006\u000eJex)'(J\u0006\u000eJ)2=U. For large tilts \u0001\u001d\n1\n2(Jex+\u000eJex)the many-body ground state are locked\nspins in an antiferromagnetic order, while for strong\nexchange coupling1\n2(Jex+\u000eJex)\u001d\u0001dimerized en-\ntangled pairs are favored. Implementing this Hamilto-\nnian requires a dynamically controllable spin-dependent\nsuperlattice [24, 25]. In the limit of isolated double\nwells\u000eJex\u0019Jex, applying a global gradient to a spin-\nindependent superlattice can locally reproduce the stag-\ngered tilts (Fig.1(a)). In this situation, a variation of\nthe parameters ( \u000eJ,\u0001) during the spin pump cycle cor-\nresponds to a modulation of ( Jex,\u0001) in the interacting\n1D spin chain. This cycle needs to be performed adiabat-\nically with respect to the intra double well exchange cou-\npling1\n2(Jex+\u000eJex). The pump cycle encircles the degen-\neracy point ( \u000eJex= 0,\u0001 = 0) as illustrated in Fig.1(a).\nDespite the antiferromagnetically ordered state being an\nexcitedmany-bodystateinthegloballytiltedsystem, the\npump can still be described by a topological invariant as\nthe pump of Eq.1 with local tilts, if the pumping is fast\ncompared to1\n2(Jex\u0000\u000eJex), which determines the gaps of\nadditional level crossings in the spectrum [26]. This can\nbe readily achieved in the experiment by choosing lattice\ndepths, for which inter-double well coupling is quenched.\nThe experimental setup consists of a 3D optical lat-\ntice with a superlattice along the x-axis and deep trans-\nverse lattices along yandzto create an array of decou-\npled 1D systems. Each system is initially occupied by\nan antiferromagnetically ordered spin chain of up j\"i=\njF= 1;mF=\u00001iand downj#i=jF= 1;mF= +1i\n87Rb atoms [27], localized on individual sites with Jex\u0019\n\u000eJex\u00190[26]. To start the pump cycle, every second\nbarrier is decreased to transfer two neighboring spins tothe ground state of a double well. Since a large mag-\nnetic gradient \u0001\u001dJexis present along x, the up (down)\nspin stays localized on the left (right) side, shortly de-\nnoted asj\";#i. The gradient is then reversed adiabati-\ncally compared to Jex. Thereby the wavefunction follows\nthe instantaneous eigenstate and a spin current occurs\nas the two spins exchange their positions via the delocal-\nized triplet state1p\n2\u0000\nj\";#i+j#;\"i\u0001\nat\u0001 = 0(Fig.1(b)).\nAt the end of the first half cycle, individual sites are\ndecoupled by increasing the short lattice depth. Subse-\nquently,\u000eJexis inverted by flipping the dimerization and\nalso the magnetic gradient to maintain the correct local\ntilts (insets Fig.1(a)). This corresponds to a projection\non double wells shifted by one lattice site. The state of\nthe system remains unchanged during this sudden switch\nand the spins in the new shifted double well are in the\nground state. After a full cycle, the two spin components\nhave each moved by 2dsin opposite directions; therefore\nthe total particle current vanishes as the contributions\nRLRφ = π\nB´x B´xU∆Jex J\nspin transportφδJex\n∆\nRLRφ = 0\nba\nds\nFigure 1. Spin pump cycle. (a) Spin pump cycle in parameter\nspace (green) of spin-dependent tilt \u0001and exchange coupling\ndimerization \u000eJex. The path can be parametrized by the an-\ngle\u001e, the pump parameter. The insets in the quadrants show\nthe local mapping of globally tilted double wells to the cor-\nresponding local superlattice tilts with the black rectangles\nindicating the decoupled double wells. Between \u001e= 0and\n\u0019,j\"iandj#ispins exchange their position, which can be\nobserved by site-resolved band mapping images detecting the\nspin occupation on the left (L) and right (R) sites, respec-\ntively. (b) Evolution of the two-particle ground state in a\ndouble well around \u0001 = 0with tunnel coupling1\n2(J+\u000eJ),\non-site interaction energy U, and spin-dependent tilt \u0001as\nwell as the exchange coupling Jex'(J+\u000eJ)2=Uand the\nlattice constant ds.3\n-500 -250 0 250 500\n∆s/h (Hz)0.000.060.120.18A\nII\nIII\nI\nImbalance \n0.60.70.8\n0 1 2\nt´ (ms)III-0.10.00.10 2.5 5\nII\n0 3 6-0.5-0.4-0.3 I\n-10 0 10-0.800.8∆/h (kHz)\n-10 0 1000.060.12a2\n-10 0 10\nt (ms)-101b a\nFigure 2. Spin current measurement. (a) Illustration of the measurement scheme. \u0001is ramped with a rate of _\u0001 = 82(2) kHz/s\nat\u0001 = 0and the ramp is stopped abruptly at different points in the cycle \u0001s(upper graph). During this ramp, the two-particle\nwavefunction initially in the ground state has a small admixture a2of the first excited state around \u0001 = 0(middle graph).\nAfter the stop of the ramp, this admixture leads to spin imbalance oscillations with an amplitude proportional to a2and thus to\nthe instantaneous spin current at ts. The lower graph shows a numerical simulation of the spin imbalance time traces assuming\nperfect adiabaticity. (b) Spin imbalance oscillation amplitude Aat different points in the pump cycle for Jex=h= 342(2) Hz\n(blue) and Jex=h= 467(3) Hz (orange). Each point is the amplitude obtained by fitting Eq.3 to the spin imbalance that was\nmeasured as a function of the holdtime t0; error bars are the fit uncertainty. Three traces are shown on the right hand side\nfor\u0001s=h=\u0000144(7)Hz (I, dark blue), \u0001s=h= 18(5)Hz (II, blue), and \u0001s=h= 530(30) Hz (III, light blue) corresponding to\nthe illustrations in (a). Each trace consists of 26 points, which were averaged five times. The light solid lines in the main\nplot show the numerical calculation for the oscillation amplitude taking into account the reduced detection efficiency due to a\nresidual exponential decay of \u0001. The dark solid lines include additionally a finite ground state occupation of 97(1) %and a\npump efficiency of 89(1) %, which were measured separately by band mapping.\nfrom the two spin components cancel each other exactly.\nThus, pumping leads to a spin transport without induc-\ning a particle current.\nThe spin current jbetween the left (L) and the right\n(R) siteof adouble well isrelated tothe change in theex-\npectation value of the spin imbalance I=1\n2(nL#\u0000nL\"\u0000\nnR#+nR\")given by the integral form of the continuity\nequation 2j=@tI, withni\u001bthe occupation of spin \u001b\non sitei. To understand how this spin current arises\nand how it can be detected, it is useful to examine the\nevolution of the eigenstates during the adiabatic change\nof\u0001(t). Two spins initially at time tiin the eigenstate\njntiiofthetwo-particledoublewellHamiltonian ^HDW(ti)\n[26] follow the instantaneous eigenstate jnti, but acquire\n– even for a perfect adiabatic evolution – a small imag-\ninary contribution iam(t)from other eigenstates jmti.\nThis admixture occurs only temporarily during the ramp\nand induces an anomalous spin velocity (Fig.2(a)). The\ncoefficients amcan be calculated in first-order perturba-\ntion theory am(t) =\u0000_\u0001hmtj~@\u0001jnti\nEn(t)\u0000Em(t)with _\u0001 =@t\u0001(t)\nthe ramp speed and El(t)the eigenenergy of jlti[28].\nWhen starting from the ground state j1ti, the wavefunc-\ntion is well approximated by only considering contribu-\ntions from the first excited state j ti\u0019j1ti+ ia2(t)j2ti.\nTheam-coefficients of higher lying states are stronglysuppressed as the corresponding wavefunctions depend\nweakly on \u0001andEm\u0000E1\u001dJex. The wavefunction j ti\ncan be probed by a sudden stop of the pump cycle at\ntimetsby projecting it onto ^HDW(ts). During the subse-\nquent time evolution, the two states j1tiandj2tiacquire\na relative phase leading to oscillations of the spin im-\nbalanceI(t) =Is+Asin[(E2(ts)\u0000E1(ts))=~\u0001(t\u0000ts)]\nwithIsthe imbalance at time ts. The oscillation am-\nplitudeA=\u00002a2(ts)h1tsj^Ij2tsiis proportional to the\nadmixture of the second eigenstate and can be related to\nthespincurrent j(ts)throughthecontinuityequation[26]\nj(ts) =AE2(ts)\u0000E1(ts)\n2~: (2)\nExperimentally, the gradient ramp was abruptly stopped\nat\u0001s= \u0001(ts)and a time trace of the resulting double\nwell superexchange oscillation [29] was recorded by a si-\nmultaneous measurement of nL\",nL#,nR\"andnR#with\nStern-Gerlach separated site-resolved band mapping im-\nages (Fig.2(a)). The amplitude Awas found by fitting\nIfit(t0) =Ae\u0000t0=\u001cexsin (!ext0+\u0012) +Is+Ide\u0000t0=\u001cd(3)\nto the oscillation data, where t0=t\u0000tsand\u0012\u00190\na phase shift induced by a finite freezing ramp speed.\nCompared to the ideal evolution, two additional effects4\nj\n250 350 450 550\nJex/h (Hz)0.00.10.2A250 350 450 550\nJex/h (Hz)0.00.51.01.5int. \nFigure 3. Imbalance oscillation amplitude Aat\u0001s= 0as a\nfunction of Jex. Each point is an average of the fitted am-\nplitudes of three time traces; the error bars are the standard\ndeviations. The blue lines are a numerical calculation taking\ninto account the reduced detection efficiency as well as the\nmeasured initial state occupation and finite pump efficiency.\nAssuming a constant scaling factor for each Jexas indicated\nby the measurement in Fig.2, the integrated current per cycle\ncan be estimated and is shown in the inset. In the ideal case,\nthe integrated current is equal to one (gray line).\nare taken into account. First, an exponential decay of\nthe amplitude with a time constant \u001cexaccounts for de-\nphasing between individual double wells. Both, \u001cexand\nthe oscillation frequency !ex'(E2\u0000E1)=~were deter-\nmined for each \u0001swith an independent superexchange\noscillation measurement. Second, an additional decay of\nthe imbalance offset Idis caused by an exponential re-\nlaxation of a small residual magnetic field gradient after\nthe abrupt stop of B0=\u000024:3(6)Hz=ds, with a decay\nconstant\u001cd= 1:05(5)ms. The resulting oscillation am-\nplitudes during the pump cycle for two different Jexas\nwell as exemplary spin imbalance traces are summarized\ninFig.2(b). Thespincurrentpeaksaround \u0001 = 0, where\nthe ground state is delocalized and spins move. For large\ngradients the eigenstates are independent of \u0001and the\nspin current vanishes. Note that the residual gradient\n\u0001dslightly shifts the peak towards higher \u0001. Further-\nmore, the spin current strongly depends on the exchange\ncouplingJex. With increasing Jex, the wavefunction de-\nlocalizes and depends less on \u0001, so the peak width in-\ncreases while the maximum amplitude decreases. How-\never, unlike the instantaneous current, the transported\nspin during one pump cycle, is independent of the pump\nparameters.\nTo compare the data with theoretical expectations, we\nperformed a numerical calculation including a reduced\ndetection efficiency caused by the residual gradient de-\ncay. Imbalance time traces for tswere evaluated using\na two spin, two-site extended Bose-Hubbard Hamiltonian\n^HDW(t0)with \u0001(t0) = \u0001de\u0000t0=\u001cd+ \u0001s[26, 30]. The cal-\nculated time traces were also fitted with Eq.3; the re-\nsulting oscillation amplitude describes ideal transport of\n0 0.5 1 1.5 2\nφ (2π )-3-2-10123x (ds )\nFigure4. Center-of-masspositionofup(red)anddown(blue)\nspins as a function of the pump parameter \u001e. The points\nshow the center-of-mass position averaged over ten data sets\nof a spin-selective imaged atom cloud; the error bars show\nthe error of the mean. Each data set consists of an average\nof ten pairs, which contain an image obtained by a sequence\nwith pumping and one using a reference sequence with the\nsame length but constant pump parameter \u001e= 0. Difference\nimages of both sequences for up and down spin are shown on\nthe right side. The solid lines depict the calculated motion\nof a localized spin for the ideal case (light gray) and for a\nreduced ground state occupation and a pump efficiency per\nhalf pump cycle that was determined independently through\na band mapping sequence (gray).\nground state spins (light lines in Fig.2). This curve can\nbe fitted to the data by rescaling the amplitude with a\nfactor of 0:84(6), which directly determines the reduction\nof the integrated measured current compared to the ideal\none and thereby the transported spin polarization. The\ndeviation from ideal transport can be attributed to an\nimperfect initial state preparation with 97(1) %ground\nstate occupation and a pump efficiency per half a pump\ncycle of 89(1) %, which describes the fraction of double\nwells that remain in the ground state after half a cycle.\nConsidering this additional occupation of the first ex-\ncited state, which creates an opposite current, the total\nspin current can be calculated (dark lines in Fig.2(b)).\nFitting these expected oscillation amplitudes to the mea-\nsured data by rescaling with a global factor results in a\nfit value of 1:05(8)forJex=h= 342(2) Hz and 1:06(8)\nforJex=h= 467(3) Hz. This shows that even though\nthe shape and amplitude of the curve changes, the in-\ntegrated current is only defined by the pumps’ topology\nnot by the specific tunneling parameters. Furthermore,\nwe note that the deviation to the theoretically expected\nintegrated current can not be attributed to edge effects\nas they are negligibly small for the present trapping po-\ntential.\nTo study the dependence of the maximum current\non the exchange coupling, the oscillation amplitude was\nmeasured at \u0001s= 0for various Jex(Fig.3). The max-\nimum amplitude decreases with rising Jex, as the spins’\nwavefunction is more delocalized for the same \u0001and5\ntherefore the current flow is spread over a larger sector\nin the pump cycle. The observed peak amplitude agrees\nwith the numerical model including the initial ground\nstate occupation and pump efficiency. As suggested by\nthe measurements in Fig.2, the integrated spin current\ncan be extracted by rescaling the ideal amplitude with a\nglobal factor and is found to be constant for all exchange\ncouplings (inset Fig.3).\nIndependent evidence for the spin separation and a\nquantitative comparison with the total spin current can\nbe obtained by measuring the center-of-mass position of\nthe two spin components from in-situ absorption images\nafter removing one of the spin components. When vary-\ning\u001e, the up and down spins clearly separate (Fig.4).\nThis independently verifies the spin transport and shows\nquantitative agreement with the results of the spin cur-\nrent measurement [26].\nIn conclusion, we have demonstrated the implemen-\ntation of a spin pump and introduced a new method for\ndirectly measuring instantaneous spin currents. Compar-\ning the measured spin imbalance oscillation amplitudes\nwith the adiabatic theory shows that the integrated cur-\nrent is independent of the specific pump parameters and\ngives evidence for the utility of the developed current\nmeasurement method. The method can also be extended\nto more general systems by performing an instantaneous\nprojection onto double wells. Investigating such spin\npumps on a single site level would allow for local obser-\nvation of spin currents and the direct observation of edge\nexcitations in finite systems [31]. A system described by\nthe non-trivial Z2-invariant can be realized with time-\nreversal invariant spin orbit interaction [9, 11]. When\nbreaking time-reversal symmetry, the topological prop-\nerties of the QSH system remain but spin-Chern num-\nbers are required for the description [11]. Furthermore,\na topological, interaction-driven quantum motor [32, 33]\ncan be accomplished by only pumping one of the compo-\nnents while the other is coupled by interaction. For spin\npumps with highly degenerate many-body ground states,\nfractional transport is predicted [34].\nWe acknowledge insightful discussions with M. Aidels-\nburger. This work was supported by NIM, the EU\n(UQUAM, SIQS), and the DFG (DIP & FOR2414).\nM.L. was additionally supported by ExQM and R.C. by\nFIRB-2012-HybridNanoDev(GrantNo. RBFR1236VV).\n[1] K. v. Klitzing, G. Dorda, and M. Pepper, Phys. Rev.\nLett.45, 494 (1980).[2] D. C. Tsui, H. L. Stormer, and A. C. Gossard, Phys. Rev.\nLett.48, 1559 (1982).\n[3] D. J. Thouless, M. Kohmoto, M. P. Nightingale, and\nM. den Nijs, Phys. Rev. Lett. 49, 405 (1982).\n[4] D. J. Thouless, Phys. Rev. B 27, 6083 (1983).\n[5] Q. Niu and D. J. Thouless, J. Phys. A 17, 2453 (1984).\n[6] M. König, S. Wiedmann, C. Brüne, et al., Science 318,\n766 (2007).\n[7] C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 226801\n(2005).\n[8] B. A. Bernevig and S.-C. Zhang, Phys. Rev. Lett. 96,\n106802 (2006).\n[9] R. Shindou, J. Phys. Soc. Jpn. 74, 1214 (2005).\n[10] L. Fu and C. L. Kane, Phys. Rev. B 74, 195312 (2006).\n[11] C. Q. Zhou, Y. F. Zhang, L. Sheng, et al., Phys. Rev. B\n90, 085133 (2014).\n[12] G. A. Prinz, Science 282, 1660 (1998).\n[13] S. Murakami, N. Nagaosa, and S.-C. Zhang, Science 301,\n1348 (2003).\n[14] Y. K. Kato, R. C. Myers, A. C. Gossard, and D. D.\nAwschalom, Science 306, 1910 (2004).\n[15] P. Sharma and C. Chamon, Phys. Rev. Lett. 87, 096401\n(2001).\n[16] R.Citro, N.Andrei,andQ.Niu,Phys.Rev.B 68,165312\n(2003).\n[17] R. Citro, F. Romeo, and N. Andrei, Phys. Rev. B 84,\n161301 (2011).\n[18] S. K. Watson, R. M. Potok, C. M. Marcus, and V. Uman-\nsky, Phys. Rev. Lett. 91, 258301 (2003).\n[19] C.W.Sandweg,Y.Kajiwara,A.V.Chumak,etal.,Phys.\nRev. Lett. 106, 216601 (2011).\n[20] M. Lohse, C. Schweizer, O. Zilberberg, M. Aidelsburger,\nand I. Bloch, Nat. Phys. 12, 350 (2016).\n[21] H.-I. Lu, M. Schemmer, L. M. Aycock, et al., Phys. Rev.\nLett.116, 200402 (2016).\n[22] S. Nakajima, T. Tomita, S. Taie, et al., Nat Phys 12, 296\n(2016).\n[23] M. J. Rice and E. J. Mele, Phys. Rev. Lett. 49, 1455\n(1982).\n[24] P. J. Lee, M. Anderlini, B. L. Brown, et al., Phys. Rev.\nLett.99, 020402 (2007).\n[25] H.-N. Dai, B. Yang, A. Reingruber, et al., Nat. Phys. 12,\n783 (2016), article.\n[26] See Supplemental Material.\n[27] A. Widera, F. Gerbier, S. Fölling, et al., Phys. Rev. Lett.\n95, 190405 (2005).\n[28] D. Xiao, M.-C. Chang, and Q. Niu, Rev. Mod. Phys. 82,\n1959 (2010).\n[29] S. Trotzky, P. Cheinet, S. Fölling, et al., Science 319, 295\n(2008).\n[30] V. W. Scarola and S. Das Sarma, Phys. Rev. Lett. 95,\n033003 (2005).\n[31] Y. Hatsugai and T. Fukui, Phys. Rev. B 94, 041102\n(2016).\n[32] F. Zhou, Phys. Rev. B 70, 125321 (2004).\n[33] R. Bustos-Marún, G. Refael, and F. von Oppen, Phys.\nRev. Lett. 111, 060802 (2013).\n[34] D. Meidan, T. Micklitz, and P. W. Brouwer, Phys. Rev.\nB84, 075325 (2011).6\nSupplemental Material for:\nSpin pumping and measurement of spin currents in optical superlattices\nC. Schweizer1;2, M. Lohse1;2, R. Citro3;4, I. Bloch1;2\n1Fakultät für Physik, Ludwig-Maximilians-Universität, Schellingstrasse 4, D-80799 München, Germany\n2Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Strasse 1, D-85748 Garching, Germany\n3Dipartimento di Fisica “E. R. Caianiello”, Università degli Studi di Salerno,\nVia Giovanni Paolo II 132, I-84084 Fisciano (Salerno), Italy\n4SPIN-CNR Salerno, Via Giovanni Paolo II 132, I-84084 Fisciano (Salerno), Italy\nSPIN CURRENT MEASUREMENT\nTwo-site extended Bose-Hubbard model\nIn the tight binding limit, two spins on an isolated\ndoublewell( \u000eJ=J)canbedescribedbyatwo-siteBose-\nHubbard model, where we denote the spins \u001b=f\";#g\nand the left (right) site with L (R). The Hamiltonian is\ngiven by\n^HBH=\u0000JX\n\u001b=f\";#g\u0010\n^ay\nL;\u001b^aR;\u001b+h.c.\u0011\n\u0000\u0001\n2(^nL;#\u0000^nL;\"\u0000^nR;#+ ^nR;\")\n+U(^nL;\"^nL;#+ ^nR;\"^nR;#)(S.1)\nwith ^ay\nR/L;\u001b(^aR/L;\u001b) the creation (annihilation) operator\nof spin\u001bon the left or right site, ^nR/L;\u001bthe number op-\nerator counting the spins, Jthe tunneling rate, \u0001the\nspin-dependent energy offset between left and right site,\nandUthe on-site interaction energy. The accuracy of\nthe exchange coupling, especially at large J=U, can be in-\ncreased by including corrections from density-dependent\nhoppingJddh=\u0000gR\nw3\nL(r)wR(r)d3rand nearest neigh-\nbor interaction ULR=gR\nw2\nL(r)w2\nR(r)d3r[S1, S2]. Here,\nwL/R(r)denotes the Wannier function on the left/right\nsite. The Hamiltonian of this extended Bose-Hubbard\nmodel ^HDWis in the basisj\"#;0i,j\";#i,j#;\"i,j0;\"#i\nwithJ0=J+Jddh:\n^HDW=0\nBB@U\u0000J0\u0000J0ULR\n\u0000J0ULR+ \u0001ULR\u0000J0\n\u0000J0ULRULR\u0000\u0001\u0000J0\nULR\u0000J0\u0000J0U1\nCCA:(S.2)\nThe dependence of the energy spectrum Enof the ex-\ntended Bose-Hubbard model on \u0001is shown in Fig.S1(a).\nWe define the exchange coupling Jexas the gap between\nthe first and second eigenenergy at \u0001 = 0. The ground\nstate in the limitj\u0001j\u001dJexisj1i\u0019j#;\"ifor positive and\nj\";#ifor negative \u0001. At \u0001 = 0andU\u001dJthe ground\nstate is approximately a triplet state1p\n2\u0000\nj\";#i+j#;\"i\u0001\n.\nNote that the third eigenstate is independent of \u0001.\n-500 -250 0 250 500\n∆/h (Hz)-0.100.1\n-500 -250 0 250 500\n∆/h (Hz)04080120\nam (10-3)am (10-3)\na2a4\na3-4-2E/h (kHz)024\n2 1\n∆/h (kHz)0 -1 -2 ba\n|1/angbracketright|2/angbracketright|3/angbracketright|4/angbracketrightFigure S1. Energy spectrum and am-coefficients. (a) Energy\nspectrum of the extended Bose-Hubbard model for two spins\non a double well. (b) Magnitude of the am-coefficients at\ndifferent spin-dependent tilt values \u0001during the magnetic\nfield ramp with a speed of _\u0001 = 82(2) kHz/s and exchange\ncoupling of Jex=h= 342(2) Hz.\nAn adiabatic change of the tilt \u0001with rate _\u0001leads in\nfirst-order approximation to the temporary wavefunction\nj ti=jnti+ iX\nm6=nam(t)jmti (S.3)\nusing parallel transport conditions [S3]. Here, jltiis an\ninstantaneous eigenstate of ^HDWand the admixture co-\nefficients are\nam(t) =\u0000hmtj~@tjnti\nEn(t)\u0000Em(t)=\u0000_\u0001hmtj~@\u0001jnti\nEn(t)\u0000Em(t):(S.4)\nA detailed derivation can be e.g. found in Ref.[S3]. The\nam-coefficients forjnti=j1ti, i.e. starting from the\nground state, are depicted in Fig.S1(b); the contribu-\ntiona2clearly dominates and other coefficients can be7\nneglected because the dependence of the eigenstates j4i\nandj3ion\u0001is weak or even vanishing because of the\nlarge gapjEm\u0000E1j\u0018U\u001dJex, form > 2. Then the\ntemporary wavefunction Eq.S.3 reduces to\nj ti\u0019j1ti+ ia2(t)j2ti: (S.5)\nConnection Between Amplitude and Current\nFor the current measurement, the pump cycle is\nabruptly stopped at \u0001s. In the ideal case this hap-\npensinstantaneously, suchthatthewavefunction j (\u0001s)i\nsubsequently evolves according to the static Hamilto-\nnian ^HDW(\u0001s). Sincej (\u0001s)iis in general not an eigen-\nstateof ^HDW(\u0001s), theexpectationvalueoftheimbalance\noperatorIoscillates in time\nI(t0) =Asin\u0012E2(ts)\u0000E1(ts)\n~t0\u0013\n+Is(S.6)\nwitht0=t\u0000ts. The oscillation amplitude is directly\nproportional to the a2-coefficient\nA=\u00002a2(ts)h1tsj^Ij2tsi: (S.7)\nWhen comparing this result to the the integral form of\nthe continuity equation, the spin current at tsis directly\nconnected to the oscillation amplitude Avia\nj(ts) =1\n2@tI(t0)\f\f\f\f\nt=ts\n=AE2(ts)\u0000E1(ts)\n2~cos\u0012E2\u0000E1\n~t0\u0013\f\f\f\f\nt=ts\n=AE2(ts)\u0000E1(ts)\n2~:(S.8)\nOscillation Amplitude – Model and Corrections\nFor two spins occupying the ground state of an isolated\ndoublewell, theidealoscillationamplitudefora perfectly\nabrupt stop of the ramp can be evaluated from Eq.S.7.\nThis ideal amplitude, within ^HDW, is shown as a func-\ntion of \u0001sin Fig.S2 (green solid lines). As the abrupt\nstop is smoothed by a decaying residual magnetic gra-\ndient, the measured amplitude for a given spin current\nis slightly reduced. This amplitude can be calculated\nnumerically by solving the time-dependent Schrödinger\nequation including the residual magnetic gradient de-\ncay. The model assumes an initial state after the adia-\nbatic evolution (Eq.S.5) stopped at \u0001s+ \u0001dand a time-\ndependent Hamiltonian ^HDW(\u0001de\u0000t0=\u001cd+ \u0001s). For long\ntimest0\u001d\u001cd, theHamiltonianapproachestheidealcase.\nWith a numerical time propagation, the evolution of the\nwavefunction and thereby I(t0)can be calculated. From\nthis time trace the amplitude is extracted with a fit of\n0.070.140.21AJex/h=467Hz\n0.00\n-500 -250 0 250 500\n∆/h (Hz)0.000.070.140.21AJex/h=342HzFigure S2. Current measurement. The figure summarizes dif-\nferent theory curves for both data sets at Jex=h= 467(3) Hz\n(orange) and Jex=h= 342(2) Hz (blue) from 2 in the main\ntext. The green solid line shows the ideally expected oscilla-\ntion amplitude. A corrected description taking into account\nthe reduced detection efficiency due to a residual exponential\ndecay of the magnetic field gradient (orange solid line), and\na description including additionally a finite ground state oc-\ncupation of 97(1) %and a pump efficiency of 89(1) %(blue\nsolid line) are shown. Furthermore, a measurement of the\nparticle current is shown (gray), obtained by analyzing the\nparticle imbalance oscillations between the left and right side\nof the double well based on the same data sets as for the spin\nimbalance oscillations.\n3 from the main text as for the experimental data. The\neffect of this detection correction is visualized in Fig.S2\nas orange solid lines.\nPreparing only ground state double wells is extremely\nchallenging and therefore a small residual amount of ex-\ncited state double wells is expected. When considering\nonly ground and first excited state, the spin imbalance is\na good measure for the state occupation at j\u0001j\u001dJex,\nwhere the spins are perfectly localized. By measuring\nthe spin imbalance before ( Ii) and after half a pump cy-\ncle (If), the initial state preparation and the pump ef-\nficiency can be characterized. The pump efficiency \f1\nis a measure of the fraction of double wells that remain\nin the ground state during half a pump cycle and can\nbe deduced from the spin imbalance: \f1=Ii\u0000If\n2Ii. The\ninitial fraction of ground and excited state double wells\nisn(i)\n1=I1+Ii\n2I1andn(i)\n2=I2+Ii\n2I2=I1\u0000Ii\n2I1, assuming\nI1=\u0000I2being the ideal spin imbalance at the initial\nconditions.8\nThe integrated spin current j=Rtf\ntij(t)dtis the sum of\nthe ideal integrated spin currents j1=2,idealof both states\nweighted by the initial band occupation and the pump\nefficiency. The ideal spin currents are oppositely directed\nj1,ideal =\u0000j2,ideal =I1and given by the ideal initial spin\nimbalance. Furthermore, equal pump efficiency \f2=\f1\nfor ground and first excited state are assumed.\nj=n(i)\n1\f1j1,ideal +n(i)\n2\f2j2,ideal\n=\u0010\nn(i)\n1\u0000n(i)\n2\u0011\n\f1j1,ideal\n=Ii\u0000If\n2(S.9)\nThe result can be verified by a comparison with the in-\ntegral form of the continuity equation 2j=@tI.\nThe reduction of the integrated current as a result of\na finite excited state occupation and a reduced pump ef-\nficiency can be approximately captured in the data anal-\nysis by rescaling of the current j(t)with a global factor.\nSuch a rescaling corresponds to a description by an aver-\nage state occupation with perfect pump efficiency. The\nblue line in Fig.S2 shows the spin current taking into\naccount the detection efficiency, the initial ground state\noccupation as well as the pump efficiency.\nENERGY SPECTRUM\nIn the experiment, the spin chain with local tilts is re-\nalized in the limit of isolated double wells with a global\nmagnetic gradient. In this system unlike for ^Hin 1\nof the main text, the prepared initial state is not the\nground state but an excited state. During the pump cy-\ncle, however, it evolves in the same way as the ground\nstate of a spin-dependent superlattice with the excep-\ntion of a number of additional crossings that occur in\nthe energy spectrum (see Fig.S3). However, the gaps are\nvery small – on the order of the inter double well ex-\nchange coupling1\n2(Jex\u0000\u000eJex)– and can be crossed non-\nadiabatically. In conclusion the gradient model requires\nnot only adiabaticity with respect to the intra double\nwell exchange tunneling1\n2(Jex+\u000eJex)but also pure non-\nadiabatictransferwithrespecttogapsontheorderofthe\ninter double well exchange coupling1\n2(Jex\u0000\u000eJex). As the\nlatter can be suppressed exponentially with the long lat-\ntice depth, and adiabaticity on the double well scale can\nbe reached by slower ramp speed, the transport in these\nmodels can be described by the same topological invari-\nant. An energy spectrum for two up and two down spins\non two double wells is shown in Fig.S3 for the experi-\nmental parameter set for Jex=h= 342(2) Hz. The pump\ncycle follows the thicker darker depicted state, which is\neither strongly gapped around \u0001 = 0or crosses states\nwith negligibly small gaps. Additionally, the state over-\nlap between the groundstate in the staggered model and\n1\n2(J\n+δJ\n)|/angbracketleft1\n|Ψ\n /angbracketright|2\n01\n-3 -2 -1 0 1 2 3\n∆/h (kHz)-4-3-2-10123E/h (kHz)\n1\n2(J\n−δJ\n)≈0Figure S3. Energy spectrum and state overlap for a superlat-\ntice with staggered tilt and a global gradient, respectively. In\nthe upper panel the energy spectrum of two up and two down\nspins on two double wells with local, staggered tilt (red) and\nglobal gradient (blue) is shown. The thicker darker lines rep-\nresent the ground state of the staggered superlattice j1stiand\nthe corresponding state in the globally tilted lattice j\u001egradi,\nwhich is used for pumping in the experiment. In the lower\npanel the state overlap between these two states is depicted.\nthe pumped state for the model with a global gradient is\ncalculated and found to be one, apart from the vicinity\nof the tiny gaps. This shows that the pumped state in\nthe gradient model is very similar to the one in the model\nwith staggered tilts.\nEXPERIMENTAL SEQUENCE\nThe experimental sequence starts with a Mott insu-\nlator of (F= 1;mF=\u00001)87Rb atoms in a 3D opti-\ncal lattice of three mutually orthogonal standing waves\nwith wavelengths \u0015x=\u0015y= 767nm and\u0015z= 844nm.\nWith a sequence of microwave driven adiabatic passages,\nthe atoms are transferred to the (F= 1;mF= 0)via\nthe(F= 2;mF= +1)state (seeft1,ft2in Fig.S4).\nThen two neighboring lattice sites are merged along the\nx-direction into decoupled sites with twice the period\nby ramping up a long lattice with period dl= 2ds,\nwhereds=\u0015x=2, and simultaneously turning the short\nperiod lattice off. With coherent microwave-mediated\nspin changing collisions [S4] each atom pair is trans-\nferred to a pair of j\"i=jF= 1;mF=\u00001iandj#i=\njF= 1;mF= +1iatoms. Subsequently, a magnetic field\ngradient is turned on and the long lattice sites are split\nadiabatically into two decoupled sites by turning on the\nshort period lattice to a depth of Vs= 40Er,sin50ms\nwithEr,i=h2\n2mRb\u00152\ni, wheremRbis the mass of a Rubid-9\n20ms60ms38.5ms38.5ms\n537 540 460 460100\nπ50\n0MWBQuad\nfsc ft2 ft1Bx\nMWBQuadBxVi (Er,i )Vz Vy\nVx,lVx,s150ms50ms50ms150ms90ms\nt (ms)\nt (ms)460 310 150 0100\n50\n0Vi (Er,i )a Initial state preparation\nb Current measurement Pump cyclec\nFigure S4. Summary of the experimental sequences. The lat-\ntice depths are shown in each upper panel in their respective\nrecoil energies: along the x-direction with period ds(blue)\nanddl(red), along the y-direction with period ds(green)\nand along the z-direction with period dz(yellow). In the\nlower panel the two contributions to the magnetic field gra-\ndient as well as the mircowave pulses are sketched. (a) A\ntime sequence for the initial state preparation is depicted. It\nends with decoupled sites occupied by antiferromagnetically\norderedj\"i=jF= 1;mF=\u00001iandj#i=jF= 1;mF= +1i\nspins. (b) Example sequence for the current measurement\nat\u0001s= 0which directly follows the initial state prepara-\ntion. (c) The spin pump sequences for the in-situ and band\nmapping measurement starting directly after the initial state\npreparation. Noteafterhalfacyclethedimerizationisflipped\nby swapping to a second laser creating \u0019shifted double wells,\nand the magnetic field gradient is reversed in 2ms.\nium atom. Due to the present magnetic field gradient\nof\u0001=h= 2:7(2)kHz during the splitting, the up and\ndown spins order antiferromagnetically in individual, de-\ncoupled sites with Jex\u00190. A detailed time sequence\nof the individual experimental parameters is shown in\nFig.S4(a). This initial state preparation is then followed\nby the pumping sequence.The spin pump sequence for multiple cycles, which\nwas used for the in-situ measurements, starts by cou-\npling neighboring lattice sites of each double well by de-\ncreasing the short period lattice to the final value for the\ncorresponding superexchange coupling. Then, the spin-\ndependent gradient is inverted adiabatically by changing\nthe bias field along the x-direction (see gradient calibra-\ntion), and at the end of the first half pump cycle the\ndistribution is frozen again by increasing the short lat-\ntice depth such that Jex\u00190. To continue the cycle, the\ndimerization is changed by swapping to a second laser\nwhich creates a \u0019-shifted long lattice, and simultane-\nously setting the magnetic bias field back to its initial\nvalue in 2ms. Then the previously described cycle is re-\npeated multiple times. Before in-situ imaging, residual\natoms in the ( F= 1,mF= 0) state are removed by ap-\nplying simultaneously a microwave field on the ( F= 1,\nmF= 0)!(F= 2,mF= 0) transition and a resonant\nimaging light pulse to remove the atoms in the F= 2\nmanifold. Subsequently, one of the spin states is selected\nand transferred to the ( F= 2,mF= 0) state by a mi-\ncrowave driven adiabatic passage and imaged by absorp-\ntion imaging.\nThe current measurement sequence is limited to the\nfirst half pump cycle. Neighboring sites are coupled\nby decreasing the short lattice depth Vsand the bias\nfieldBxis changed with a constant rate to different fi-\nnal values of the magnetic tilt \u0001s, where it is abruptly\nstopped. After a variable holdtime t, the spin distribu-\ntionisfrozenbyrampinguptheshortperiodlatticeanda\nsite-resolved band mapping was performed. During time-\nof-flight, the spin-components are separated spatially by\na Stern-Gerlach field and the four site occupations nL#,\nnL\",nR#, andnR\"can be simultaneously extracted from\neach absorption image.\nCALIBRATION OF LATTICE, GRADIENT\nFIELD AND HUBBARD PARAMETERS\nCalibration of bare tunneling rate: The bare tunnel-\ning matrix element Jcouples two neighboring sites and\ncan be calibrated from left-right oscillations of a single\natom on a double well with degenerate on-site energies\n(\u0001 = 0). At the beginning, a single atom is prepared\non the left site of each double well. To this end, an\nn= 1Mott insulator in the long lattice is prepared and\nadiabatically split with the short lattice in the presence\nof a large potential tilt. The short lattice depth is in-\ncreased such that the atoms localize on the lower-lying\nleft site (J\u00190). Then, the tilt is removed and subse-\nquently the short period lattice is ramped down in 200\u0016s\nto the lattice depth at which Jneeds to be calibrated.\nThe localized wavefunction is not an eigenstate of the\nsymmetric double well, but an equal superposition of the\ngroundj1iand first excited state j2iand thus the left-10\n-0.8 -0.4 0.0 0.4 0.8\nBx (G)-1.2-0.60.00.61.2∆/h (kHz)\nFigure S5. Calibration of the magnetic gradient. The data\npoints show a measurement of the magnetic tilt \u0001by laser-\nassisted tunneling spectroscopy in a tilted double well as a\nfunction of the magnetic bias field Bx. From the measured\nfrequency, the tilt can be directly inferred by 2\u0001 =hfr\u0000\u000e0.\nThe error bars show the standard deviation of an average of\neight points, four for ( F= 2,mF=\u00062) each.\nrightoccupationoscillatesintime. Theleft-rightfraction\ncan be measured with site-resolved band mapping and its\noscillation frequency fbareis equal to the difference of the\nsingle-particle eigenenergies hfbare=\u000f2\u0000\u000f1= 2J.\nCalibration of on-site interaction energy: The on-site\ninteraction energy Uis the extra energy for placing a\nsecond particle on the same site and can be calibrated\nvia the superexchange oscillation frequency. An up and\na down spin are localized on the left and right site\nof a double well prepared in the presence of a large\nspin-dependent tilt \u0001as for the current measurement.\nThen, the magnetic fields, except for a small bias field to\nmaintain the quantization axis, are switched off to non-\nadiabatically remove the tilt while Jex\u00190and after-\nwards the short period lattice is decreased within 200\u0016s\nto the final value, at which Uis calibrated. The local-\nized state is not an eigenstate anymore and evolves in\ntime. The time evolution leads to superexchange oscilla-\ntions [S2], which for the experimentally used parameters\nare dominated by the contributions from the ground and\nfirst excited state ~!ex'E2\u0000E1=Jex. The exchange\ncouplingJexcan be calculated from an extended Bose-\nHubbard model and depends strongly on U. Thus,Ucan\nbe inferred from !ex.\nGradient Calibration: The spin-dependent gradient\nfield is generated by two pairs of coils: an anti-Helmholtz\npair along the z-direction, which creates a quadrupole\nfieldBquad(xex+yey\u00002zez), and a Helmholtz pair\nalong thex-direction, which creates a homogenous bias\nfieldBxexwitheithe unit vector in i-direction. At\nthe atom cloud ( x\u00190), the total magnetic field is\nB=p\n(Bquadx+Bx)2+B2\nofs(y;z)and the magnetic\nfield gradient B0=@B=@xdepends linearly on the bias\nfieldBxforBx\u001cBofs. A precise knowledge of the\n0 2 4 6\nt´ (ms)∆/h (Hz)\n05101520\n-2 0 2 4 6 8 10\nt´ (ms)∆/h (Hz)\n050100150200Figure S6. Calibration of the residual exponential decay of \u0001\nafter the ramp stop t0=t\u0000ts. The data points show the\nmagnetic gradient determined by microwave spectroscopy on\nthe (F= 1,mF=\u00001)!(F= 2,mF=\u00002) transition when\nstopping the gradient ramp at \u0001s= 0. The linear ramp be-\nfore and the exponential decay after the stop time tsis clearly\nvisible. By fitting a linear ( t <0) and an exponential func-\ntion (t >0) to the data, the decay constant \u001cd= 1:05(5)ms\nand\u0001d, which depends on the double well site distance, can\nbe determined.\nspin-dependent double well tilt is required for the current\nmeasurementmethod. Tothisend, acalibrationwasper-\nformed using laser-assisted tunneling spectroscopy. The\nsetup comprises two beams interfered under an angle of\n90\u000ewith a frequency difference of \u000e!forming a running\nwave lattice oriented at 45\u000eto the physical lattice. A\nsingle spin in the ( F= 2,mF=\u00062)-state is loaded in\nthe ground state of a tilted double well potential with\n\u000e0=h= 4:91(2)kHz and additional magnetic tilt 2\u0001,\nwhich needs to be calibrated. The running wave lattice\nmodulates neighboring lattice sites relative to each other\nand will induce tunneling if ~\u000e!= 2\u0001 +\u000e0. A series of\nspectroscopy scans varying \u000e!are performed for various\nmagnetic field values Bx. The resulting data is shown in\nFig.S5. Forlarge Bxa deviationfrom thelinear behavior\nis visible, which can be captured by fitting the magnetic\nfield distribution of a quadrupole field\n\u0001/BxBquadp\nB2x+B2\nofs; (S.10)\nwhere both BquadandBofsare fit variables.\nDecay Calibration: During the current measurement\nsequence, the magnetic field gradient is ramped to a final\nvalue \u0001sand abruptly stopped there. Experimentally,\nan instantaneous stop is not realizable but a small resid-\nual gradient remains which decays slowly during the cur-\nrent measurement. The calibration of the decay time \u001cd\nas well as the residual gradient \u0001dattsis essential for\nthe current measurement and is realized with microwave\nspectroscopy. The sequence starts with a single spin in\nthe (F= 1,mF=\u00001) state localized on the left site by a\nstrong magnetic gradient. The gradient is reduced with11\nthe identical rate as for the current measurement and\nstopped at \u0001s= 0but individual sites are decoupled by\na large short lattice depth. Now, resonance frequency\nscans on the ( F= 1,mF=\u00001)!(F= 2,mF=\u00002)\nmicrowave transition with a pulse duration of 44\u0016s are\nrecorded at various times around ts; the gradient \u0001(t0)\ndetermined from the center frequencies is summarized in\nFig.S6. A clear exponential decay \u0001de\u0000t0=\u001cdcan be fit-\nted fort0>0with a decay constant \u001cd= 1:05(5)ms. The\namplitude \u0001dcan be calibrated by comparing the linear\nincrease for t0<0with the gradient calibration (Fig.S5)\nand leads to \u0001d=h= 24:3(6)Hz for sites separated by ds.\nThis corresponds to a tilt in the superlattice for the ex-\nperimental parameters of \u0001d=h= 19:8(5)Hz atJex=h=\n342(2)Hz and \u0001d=h= 19:4(5)Hz atJex=h= 467(3) Hz.\nThe difference originates mainly in the slightly different\ndistance between the sites depending on the double well\nparameters.\nSimultaneous band mapping of two spins: The simul-\ntaneous site-resolved detection of two spins per double\nwell suffers from an additional reduction of the detected\nimbalance. This reduction occurs during the merging of\nthe left and right site of each double well in the pres-\nence of a spin-independent tilt into a single site of the\nlong lattice. This process transfers atoms from the left\nsite to the ground band and spins from the right site to\nthe second excited band of the long lattice. The subse-\nquent band mapping measures the band occupation and\nhence also the site occupations. However, spins in these\nbands undergo singlet-triplet oscillations after and also\nduring the merging [S2, S5]. A calibrated holdtime be-\nfore the release is chosen such that correct imbalances\nare detected. Nevertheless, the detected imbalance is re-\nduced most likely due to dephasing during the merging\nramps. The value of this reduction can be calibrated to\nrescale the measured imbalances with a constant factor\nto correctly determine the imbalance. The calibration\nmeasurement compares the two spin site-resolved band\nmapping with a single spin band mapping, where shortly\nbeforethemergingoneofthespincomponentsisremoved\nby an adiabatic spin transfer and a subsequent resonant\nlight pulse.\nMULTIPLE PUMP CYCLES\nBand mapping data\nThe in-situ data show a separation and opposite trans-\nport of the two spin components for multiple pump cy-\ncles. In Fig.S7 the spin imbalance Isduring the pump\ncycle is depicted versus the pump parameter \u001e, which is\ndefined as the angle of the pump path in parameter space\n(\u000eJex=\u000eJex,max;\u0001=\u0001max). The imbalance starts with a\nnegative value, the state is predominantly j\"#i, and in-\n0.0 0.5 1.0 1.5\nφ (2π)-1.0-0.50.00.51.0Figure S7. Static spin imbalance Iduring the pump cycle.\nThe data is shown with respect to the pump parameter \u001e\ndefined as the angle of the pump path in parameter space\n(\u000eJex=\u000eJex,max;\u0001=\u0001max). The error bars show the error of\nthe mean of five repetitions each.\nverts during the first half pump cycle. After switching\nthe dimerization, the pump cycle continues by inverting\nthe spin-imbalance each half pump cycle. At \u001e= 2\u0019a\nsmall step inIis visible, which originates mainly from\nsingly occupied sites created at the surface of the atom\ncloud during pumping.\nInitial state and pump efficiency\nFor the in-situ measurement the calculated motion of a\nlocalized spin is shown, which takes into account the ini-\ntial ground state occupation n(i)\n1and pump efficiency \fi\nperi-th half pump cycle. The model is analogous to the\none for the spin current, where the corrections are ex-\ntracted from the spin-imbalance measured at each half\npump cycle. The step height is then given by\nsi=\fi\u0010\n2n(i)\n1i\u00001Y\nj=0\fj\u00001\u0011\n(S.11)\nwithn(i)\n1= 0:94and\f0:::4=f1;0:97;0:91;0:96;0:90g. In\ntotal, the displacement after the i-th half pump cycle is\nx=X\nj=1sj: (S.12)\nPUMP SCHEME IN A TIGHT-BINDING MODEL\nWITH MAGNETIC FIELD GRADIENT\nIn the tight-binding approximation the dynamics of\nnon-interacting atoms in an optical superlattice poten-\ntial in the presence of a field gradient is described by a\ngeneralized Harper model with a site-dependent Zeeman-12\nlike term\n^Hs=^HJ+^H\u0001\n=\u0000X\nm;\u001b1\n2(J+\u000eJm) (^ay\nm+1;\u001b^am;\u001b+h.c.)\n+X\nmm\u0001\u0010\n^ay\nm;\"^am;\"\u0000^ay\nm;#^am;#\u0011(S.13)\nwith\u000eJm= (\u00001)m\u000eJ. The pumping scheme is imple-\nmented with a cycle, in which (\u000eJ;\u0001)!(\u000eJ(\u001e);\u0001(\u001e))\nand where the pump parameter changes constantly in\ntime\u001e= 2\u0019t=T. In the deep tight-binding regime the\nparameter space describes an ellipse (\u000eJ(t);\u0001(t)) =\n(\u000eJsin(2\u0019t=T );\u0001 cos(2\u0019t=T )). At \u0001 = 0, the spectrum\nhas an energy gap \u0001E= 2\u000eJ, and thus the adiabatic\ncondition is met if T\u001d~=\u000eJ.\nThe part of the Hamiltonian ^H\u0001is odd while the part\n^HJis even under time-reversal symmetry and therefore\nEq.S.13 belongs to the class that satisfy the condition\n^H[\u0000t] =^\u0002^H[t]^\u0002\u00001, where ^\u0002is the time-reversal opera-\ntor. Moreover, the Hamiltonian is time-reversal invariant\nat two points t1=T\n4andt2=3T\n4, where ^HJdominates.\nThe existence of these two points plays a crucial role in\nthe classification of the pump cycle. In particular, pump\ncycles in which ^H[t1]and ^H[t2]have different time re-\nversal polarization are topologically distinct from trivial\ncycles and define a Z2spin-pump. In a single double well\nat timet= 0,^H\u0001dominates and locks the up (down)\nspins on the left (right) well, denoted by j\";#i. This\nstate evolves into the j#;\"istate after half a pump cycle\natt=T\n2, where the two spins have exchanged their po-\nsitions. In contrast, at t=T\n4andt=3T\n4, the term ^HJ\ndominates and the spins are delocalized over the double\nwells; then the system is dimerized.\nCENTER OF MASS SHIFT AND\nTIME-REVERSAL POLARIZATION\nConsider a Hamiltonian Eq.S.13 with lattice constant\nds=dl=2 = 1and periodic boundary conditions. Then,\nin absence of spin-orbit type of interaction that means\nindependent spin components without inter-spin inter-\nactions, the spin transport for a homogeneously popu-\nlated band is characterized by the spin Chern number\nCsc=\u0017\"\u0000\u0017#. Since the spin components are decoupled\neven in the presence of a field gradient, the Chern num-\nbers\u0017\u001bcan be evaluated using the Thouless-Kohmoto-\nNightingale-Nijs expression [S6]:\n\u0017\u001b=1\n2\u0019ZT\n0dtZ\u0019\n\u0000\u0019dk\n\u001b(t;k);(S.14)\nwhere \n\u001bis the Berry curvature associated to the single-\nparticle wavefunction\n\n\u001b(t;k) = i (h@tu\u001bj@ku\u001bi\u0000h.c.): (S.15)The spin Chern number can be furthermore related for\nthe non-interacting case to the Z2topological invariant\nI=mod 2(Csc=2)thatdistinguishesanontrivial Z2pump\nfrom a trivial one and is related to the change in time\nreversal polarization.\nAs known from polarization theory [S7], the charge\npolarization is the center of mass of a localized Wan-\nnier state and is in turn related to Berry’s phase of the\ncorresponding Bloch functions. In the same way, the po-\nlarization of a single spin component is given by:\nP\u001b=1\n2\u0019Z\u0019\n\u0000\u0019dkA\u001b(k); (S.16)\nwhereA\u001b(k) = iPhu\u001bj@ku\u001biis the Berry connection.\nThe change in polarization induced by changing the\npump parameter \u001eby2\u0019, or the time variable, corre-\nsponds to the Chern number [S7]\n\u0017\u001b=Z2\u0019\n0d\u001e @\u001eP\u001b(\u001e): (S.17)\nThe spin-density can be directly measured by in-situ\nabsorption imaging for a single spin component. The\nchange of spin-polarization \u0001P\u001b=P\u001b(\u001e1)\u0000P\u001b(\u001e2)at\ntwo different times t1andt2, coincides with the spatial\nshift of the Wannier function. Measuring the center of\nmass shift for a single spin component thus gives the\nChern number of this component.\nFor time-reversal invariant systems, taking into ac-\ncount the role of Kramer’s degeneracy, one can define\na corresponding time-reversal polarization in terms of\nthe difference of the individual spin polarizations Ps=\nP\"\u0000P#. Hence, the change in time-reversal polarization\nduringacyclegivesthe Z2topologicalinvariant. Further-\nmore, it is equal to the integration of the instantaneous\nspin-current jover the pump cycle as,RT\n0dtj(t).\nSPIN PUMPING WITH INTERACTIONS\nWhen in addition hardcore interactions between the\nspin components are assumed, spin pumping can be un-\nderstood in a similar way as in the non-interacting case.\nFor half filling a representation in terms of spin operators\ncan be introduced in this limit:\n^S+\nm= ^ay\nm\"^am#;\n^S\u0000\nm= ^ay\nm#^am\";\n^Sz\nm= ^ay\nm\"^am\"\u0000^ay\nm#^am#:(S.18)13\nModel Eq.S.13 can thus be mapped to a model of an\nantiferromagnetic spin chain with two perturbations\n^Heff=^Hxy+^Hdim+^H\u0001\n=\u0000Jex\n4X\nm(^S+\nm^S\u0000\nm+1+h.c.)\n\u0000\u000eJex\n4X\nm(\u00001)m(^S+\nm^S\u0000\nm+1+h.c.)\n+ \u0001X\nmm^Sz\nm;(S.19)\nwhere the second term describes a staggered compo-\nnent of the exchange interaction, while the last one is\na Zeeman-like term, which controls the on-site energies.\nA cycle, in which (\u000eJex;\u0001)are adiabatically var-\nied defines a topological spin pump [S8]. Such a spin\npump transfers Sz=~per cycle, which can be still\ndescribed by the Z2topological invariant. Note that\nthe Hamiltonian ^Hxyin Eq.S.19 corresponds to a co-\nsine band \u000f(k) =\u0000Jex=2 cos(k). The staggered ex-\nchange interaction \u000eJexopens a gap at k=\u0006\u0019\n2and for\nhalf filling, only the lowest subband is occupied. Due\nto the\u0019periodicity in k-space, the double degenerate\npoint (k; \u000eJex;\u0001) = (\u0019\n2;0;0)is identical to that at\n(\u0000\u0019\n2;0;0)and becomes the source and sink for a vector\nfieldB+1andB\u00001defined in the k-\u001e-parameter space.\nIf a pump path \rencloses the origin (\u000eJex;\u0001) = (0;0),\ne.g. (\u000eJex;\u0001) = (\u000eJex,max cos\u001e;\u0001maxsin\u001e), where\n\u001e: 0!2\u0019, the number of lattice sites that a spin is\ntransported, is given by the flux B+1enclosed by the\npath\nI\n\rZ\u0019\nk=\u0000\u0019dS\u0001B+1= 1: (S.20)\nThis corresponds to a quantized spin transport. The to-\ntalSzat one end of this system increases while that at\nthe other end decreases by one during the entire cycle as\nlong as the gap is maintained open and the point (0;0)\nis not outside the 2D closed surface.\nAway from the hard-core constraints for the bosons,\nthe effect of a finite interaction can be taken into account\nvia a bosonization approach. When applying Haldane’s\nbosonization of interacting bosons [S9] to the Hamilto-\nnian Eq.S.13 and \u000eJex;\u0001 = 0, the Hamiltonian of the\nbosons can be written as:\n^H0=X\n\u001bZdx\n2\u0019\u0014\nv\u001bK\u001b(\u0019\u0005\u001b)2+v\u001b\nK\u001b(@x\b\u001b)2\u0015\n;(S.21)where the two canonical fields fulfill [\b\u000b(x);\u0005\f(x0)] =\ni\u000e\u000b\f\u000e(x\u0000x0),v\u001bis the velocity of excitations, and K\u001bis\nthe Tomonaga-Luttinger exponent. In the case of hard-\ncore bosons, v\u001b=Jexsin(\u0019\u001a0\n\u001b)andK\u001b= 1, while\u001a0is\nthe boson density.\nIntroducing the fields \u0012\u000b=\u0019Rx\u0005\u000b, the boson annihi-\nlation operators can be represented as [S9]:\naj\u001b= \u001b(x) (S.22)\n=ei\u0012\u001b(x)+1X\nm=0c\u001b\nmcos(2m\b\u001b(x)\u00002m\u0019\u001a(0)\n\u001bx);\nwherec\u001b\nmare non-universal coefficients. For hardcore\nbosons at half filling, these coefficients have been found\nanalytically [S10]. From Eq.S.22, the bosonized expres-\nsion of the staggered hopping term can be deduced:\n^Hhop./\u000eJZ\ndxsin(2\bc) cos(2\bs);(S.23)\nwith only the most relevant term in the renormalization\ngroup sense and the charge and spin variables \b\"=#=\n(\bc\u0006\bs). When \bcis pinned (e.g. at commensurate\nfillings) in the gapped spin phase also the field \bsis\npinnedh\bsi\u0011\u0019\n4(1 +sign(\u000eJ)). The excitations above\nthe ground state are solitons and antisolitons, which are\ntopological excitations of the field \bsthat carry a spin\n1=2.\nThe time reversal-polarization is identified as Ps=\nmod 2(2\bs\n\u0019)and because under time reversal ^\u0002\bs^\u0002\u00001=\n\u0000\bs, time reversal polarization is either 0or1. Thus,\nthe topological classification of the spin pump remains\nalso away from the hard-core bosons limit.\n[S1] V. W. Scarola and S. Das Sarma, Phys. Rev. Lett. 95,\n033003 (2005).\n[S2] S. Trotzky, P. Cheinet, S. Fölling, et al., Science 319,\n295 (2008).\n[S3] D. Xiao, M.-C. Chang, and Q. Niu, Rev. Mod. Phys.\n82, 1959 (2010).\n[S4] A. Widera, F. Gerbier, S. Fölling, et al., Phys. Rev.\nLett.95, 190405 (2005).\n[S5] M. Anderlini, P. J. Lee, B. L. Brown, et al., Nature 448,\n452 (2007).\n[S6] D. J. Thouless, M. Kohmoto, M. P. Nightingale, and\nM. den Nijs, Phys. Rev. Lett. 49, 405 (1982).\n[S7] N. Marzari and D. Vanderbilt, Phys. Rev. B 56, 12847\n(1997).\n[S8] R. Shindou, J. Phys. Soc. Jpn. 74, 1214 (2005).\n[S9] F. D. M. Haldane, Phys. Rev. Lett. 47, 1840 (1981).\n[S10] A. A. Ovchinnikov, J. Phys. Condens. Matter 16, 3147\n(2004)." }, { "title": "1005.3853v2.Cavity_spin_optodynamics.pdf", "content": "A spin optodynamics analogue of cavity optomechanics\nN. Brahms1and D.M. Stamper-Kurn1;2\u0003\n1Department of Physics, University of California, Berkeley CA 94720, USA\n2Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA\n(Dated: October 30, 2018)\nThe dynamics of a large quantum spin coupled parametrically to an optical resonator is treated in\nanalogy with the motion of a cantilever in cavity optomechanics. New spin optodynamic phenomena\nare predicted, such as cavity-spin bistability, optodynamic spin-precession frequency shifts, coherent\nampli\fcation and damping of spin, and the spin optodynamic squeezing of light.\nCavity optomechanical systems are currently being ex-\nplored with the goal of measuring and controlling me-\nchanical objects at the quantum limit, using interactions\nwith light [1]. In such systems, the position of a mechan-\nical oscillator is coupled parametrically to the frequency\nof cavity photons. A wealth of phenomena result, in-\ncluding quantum-limited measurements [2], mechanical\nresponse to photon shot noise [3], cavity cooling [4], and\nponderomotive optical squeezing [5].\nConcurrently, spins and psuedospins coupled to elec-\ntromagnetic cavities are being researched in atomic [6],\nionic [7], and nanofabricated systems [8, 9], with appli-\ncations including magnetometry [10], atomic clocks [11],\nand quantum information processing [6, 9, 12]. In con-\ntrast to mechanical objects, spin systems are more eas-\nily disconnected from their environment and prepared in\nquantum states, including squeezed states [11].\nIn this Rapid Communication, we seek to link these\ntwo \felds by exploiting the similarities between large-\nspin systems and harmonic oscillators [13] to construct a\ncavity spin optodynamics system in analogy to cavity op-\ntomechanics. Optomechanical phenomena map directly\nto our proposed system, resulting in spin cooling and\nampli\fcation [14, 15], nonlinear spin sensitivity and spin-\ncavity bistability [16, 17], and spin opto-dynamic squeez-\ning of light [14, 18]. Such a system may \fnd application\nas a quantum-limited spin ampli\fer or as a latching spin\ndetector. We detail these phenomena using currently ac-\ncessible parameters, and we propose realizations either\nusing cold atoms and visible light or using cryogenic solid\nstate systems and microwaves.\nAn ideal cavity optomechanics system, consisting of\na harmonic oscillator coupled linearly to a single-mode\ncavity \feld, obeys the Hamiltonian\nH=~!c^n+~!z^ay^a\u0000fzHO\u0000\n^ay+ ^a\u0001\n^n+Hin=out:(1)\nHere ^ais the oscillator's phonon annihilation operator,\n^nis the photon number operator, !zis the natural fre-\nquency of the oscillator in the dark, and !cis the bare\ncavity resonance frequency. fis the radiation-pressure\nforce applied by a single photon, while zHO=p\n~=2m!z\nis the harmonic oscillator length for oscillator mass m.\nHin=outdescribes the coupling of the cavity \feld to exter-\nnal light modes. Under this Hamiltonian, the cantileverposition ^zand momentum ^ pevolve asd^z=dt = ^p=m and\nd^p=dt =\u0000m!2\nz^z+f^n.\nTo construct a spin analogue of this system, we con-\nsider a Fabry-Perot cavity with its axis along k(Fig.\n1). For the collective spin, we \frst consider a gas of\nNhydrogenlike atoms in a single hyper\fne manifold of\ntheir electronic ground state, each with dimensionless\nspinsand gyromagnetic ratio \r. The atoms are op-\ntically con\fned at an antinode of the cavity \feld. An\nexternal magnetic \feld B=Bbis applied to the atoms.\nThe detuning \u0001 cabetween the cavity resonance is cho-\nsen to be large compared to both the natural linewidth\nand the hyper\fne splitting of the atoms' excited state.\nIn this limit, spontaneous emission may be ignored and\nthe single-atom cavity-\feld interaction energy, HStark =\n(~g2\n0=\u0001ca)^n(1\u0006\u001dk\u0001^ s), comprises the scalar and vector\nac Stark shifts [19], where g0quanti\fes the atom-cavity\ncoupling and \u001dthe vector shift.\nSumming over all atoms q, we obtain the system\nHamiltonian:\nH=~!c(^n++ ^n\u0000) +Hin=out +X\nq\u0012\n\u0000~\rB\u0001^ sq\n+~g2\n0\n\u0001ca[(^n++ ^n\u0000) +\u001d(^n+\u0000^n\u0000)k\u0001^ sq]\u0013\n;(2)\nwith number operators ^ n\u0006for the\u001b\u0006polarized optical\nmodes.\nThe above Hamiltonian can be rewritten as the inter-\naction of the collective spin operator ^S\u0011P\nq^ sqwith an\ne\u000bective total magnetic \feld Be\u000b\u0011\ne\u000b=\r, giving [20]\n\ne\u000b= \nLb+ \nc(^n+\u0000^n\u0000)k: (3)\nHere \nL=\rBand \nc=\u0000\u001dg2\n0=\u0001ca. Altogether, the\ncavity spin optodynamical Hamiltonian is\nH=~\u0012\n!c+Ng2\n0\n\u0001ca\u0013\n(^n++ ^n\u0000)+Hin=out\u0000~\ne\u000b\u0001^S:(4)\nNow consider the external magnetic \feld to be static\nand oriented along i, orthogonal to the cavity axis. In\nthe limith^Si'Si, the spin dynamics become\nd^Sj\ndt= \nL^Sk\u0000\ncS(^n+\u0000^n\u0000);d^Sk\ndt=\u0000\nL^Sj:(5)arXiv:1005.3853v2 [quant-ph] 3 Jan 20112\nS\nΩLiΩc (n+-n-)k\nσ+ σ -\nFIG. 1. (Color) An ensemble of atoms trapped within a driven\noptical resonator experiences an externally imposed magnetic\n\feld along iand a light-induced e\u000bective magnetic \feld along\nthe cavity axis k. The evolution of the collective spin ^Sre-\nsembles that of a cantilever in cavity optomechanics.\nThe analogy between cavity optomechanics and\nspin optodynamics is established by assigning\n^z! \u0000zHO^Sk=\u0001SSQL and ^p!pHO^Sj=\u0001SSQL,\nwherezHOandpHO =~=(2zHO) are de\fned with\n!z!\nL[13] and \u0001 SSQL =p\nS=2 is the standard\nquantum limit for transverse spin \ructuations. Eqs. (5)\nnow match the optomechanical equations of motion\nwith the optomechanical coupling de\fned through\nfzHO^n!\u0000~\nc\u0001SSQL(^n+\u0000^n\u0000). The main result of this\nwork, that various cavity optomechanical phenomena\nare manifest also in cavity spin optodynamical systems,\nis immediately established.\nLet us now elaborate on these phenomena. To obtain\ngeneral results, we will proceed without assuming ^S'Si,\nexcept in certain cases, noted in the text, where some\nphysical insight is gained. We begin with e\u000bects for which\nboth the light \feld and the ensemble spin may be treated\nclassically, i.e. by letting S=h^Siand \u0016n\u0006=h^n\u0006i.\nCavity-spin bistability: We start with the static\nbehavior of the system by \fnding the \fxed points of the\nsystem. The collective spin vector is static when Sis\nparallel to \ne\u000b. Writing S=S(isin\u00120+kcos\u00120), this\ncondition requires \u0016 n+\u0000\u0016n\u0000= (\nL=\nc) cot\u00120. The intra-\ncavity photon numbers are determined also by the stan-\ndard expression for a driven cavity of half line-width \u0014,\ni.e. \u0016n\u0006= \u0016nmax;\u0006[1 + (\u0001p;\u0006\u0006\ncScos\u00120)2=\u00142]\u00001with\n!\u0006= (!c+Ng2\n0=\u0001ca) + \u0001p;\u0006being the frequency of\nlaser light of polarization \u001b\u0006driving the cavity and\n\u0016nmax;\u0006characterizing its power. These two expressions\nfor \u0016n+\u0000\u0016n\u0000may admit several solutions (Fig. 2).\nAs typical in instances of cavity bistability [22], sev-\neral of the static solutions for the intracavity intensities\nmay be unstable. To identify such instabilities, we con-\nsider the torque on the collective spin when it is dis-\nplaced slightly toward + kfrom its static orientation.\nStable dynamics result when such displacement yields a\ntorque N\u0001jwith the sign \u000b= sgn(sin\u00120). Geometri-\ncally, this stability requires that the spin vector be dis-\nplaced further in the + kdirection than the vector \u000b\ne\u000b.\nQuantifying the linear response of the intracavity e\u000bec-\nS\n15\n0\n-10 -5 0 5 1015\n0(a)(b)\nΔp/κ Δp/κn -n+\nθ0π\n-π\n-10 -5 0 5 10FIG. 2. (Color) Cavity-spin bistability in a cavity driven\nwith linearly polarized light. We consider N= 5000 spin-\n287Rb atoms, \n c=\u0014= 1:25\u000210\u00003, \nL=\u0014= 3:3\u000210\u00002, and\n\u0016nmax;\u0006= 15 (similar to Ref. [21]). (a) As \u0001 pis varied, several\nstable (black) and unstable (gray) static spin con\fgurations\nare found. Con\fgurations for \u0001 p=\u0014=\u00004:8 are depicted. (b)\nThe cavity exhibits hysteresis as the probe is swept with posi-\ntive (dashed blue) or negative (red) frequency chirps, with the\nspin initially along i. Rapid transitions as \u0001 p=\u0014is swept up-\nward from -2.8 or downward from 0 involve symmetry break-\ning as the cavity becomes birefringent; we display \u0016 n+and \u0016n\u0000\nassuming the stable branch closer to \u00120= 0 is selected. Here,\n\u0001ca=2\u0019= 20 GHz from the D2 transition, g0=2\u0019= 15 MHz,\n\u0014=2\u0019= 1:5 MHz.\ntive magnetic \feld to variations of the collective spin via\n\u0015= \ncd(\u0016n+\u0000\u0016n\u0000)=dSk, the static spin orientations are\nfound to be unstable when \u000b\u0015> \nLjcsc3\u00120j=S.\nOpto-dynamical Larmor frequency shift: The\ndynamics of the spin precessing about one of the stable\ncon\fgurations can be parameterized by the precession\nfrequency, which is shifted from \n Lby two e\u000bects. First,\nthere is an upward frequency shift from the static modi-\n\fcation of the e\u000bective magnetic \feld, leading to preces-\nsion at the frequency \n L0= \nLjcsc\u00120jwhen\u0015= 0. A\nsecond shift occurs when the spin dynamics are slow com-\npared to the response time of the cavity \feld (\n L0\u001c\u0014).\nHere, the precessing spin modulates the cavity \feld,\nwhich, in turn, acts back upon the spin to modify its\nprecession frequency, When the precession amplitude is\nsmall, a solution of the spin equations of motion derived\nfrom the Hamiltonian in Eq. (4) yields an overall preces-\nsion frequency \n L00, where\n\nL002= \nL02\u0000\u0015\nLSsin\u00120: (6)\nThe quantity kS\u0011\u0000\u0015\nLSsin\u00120serves as the analogue\nof the optical spring constant [23], and leads to shifts of\nthe Larmor precession frequency with a sign and mag-\nnitude that depend on the spin orientation, \u0015and the\nfrequency, intensity, and polarization of the cavity probe\n\felds. When the precession amplitude is large, the dy-\nnamics become essentially nonlinear. In this case the dy-\nnamics can be described by numerical simulation (Fig. 3).\nCoherent ampli\fcation and damping of spin:3\nNow we consider the e\u000bects of the \fnite cavity response\ntime\u0014\u00001on the spin dynamics. To develop an intu-\nitive picture, we consider the unresolved sideband regime\n\nL<\u0014, in a frame (indicated by the index \\r\") corotat-\ning with the collective spin, with iraligned to the \fxed\npoint. We assume the spin to be precessing at a near\nconstant rate, and the cavity \feld response to this pre-\ncession to be simply delayed by \u0014\u00001. Employing the\nrotating-wave approximation, the delay causes the ef-\nfective \feld \ne\u000b,rto point out of the ir-krplane, with\n\ne\u000b,r\u0001jr=\u0000(\u000b\u0015Sk;rsin2\u00120sin\u001e)=2, where\u001e= \nL00=\u0014.\nThe collective spin now experiences a torque in the kr\ndirection, giving\ndSk;r\ndt=\u0000\u000b\u0015sin2\u00120sin\u001eSi;r\n2Sk;r (7)\nFor positive (negative) values of \u000b\u0015, the Larmor pre-\ncession frequency is shifted down (up) and the spin is\ndamped toward (ampli\fed away from) its stable point.\nSimilar relations apply to cavity optomechanics [24]. The\nde\rection of the spin toward or away from the stable\npoints persists for large precession amplitudes (Fig. 3).\nThis cavity-induced spin ampli\fcation or damping dif-\nfers from conventional optical pumping in two important\nrespects. First, while the spin polarization generated by\noptical pumping relies on the polarization of the pump\nlight, the target state for cavity-induced spin damping\nis selected energetically. Similar to cavity optomechani-\ncal cooling [4], cavity enhancement of Raman scattered\nlight drives spins to the high- or low-energy spin state\naccording to the detuning of probe light from the cavity\nresonance, independent of the polarization. Second, this\nampli\fcation or damping of the intracavity spin is coher-\nent, preserving the phase of Larmor precession, at least\nwithin the limits of a quantum ampli\fer.\nSpin optodynamical squeezing of light: We now\nconsider quantum optical e\u000bects of cavity spin optody-\nnamics. One such e\u000bect is the disturbance of the collec-\ntive spin due to quantum optical \ructuations of the cav-\nity \felds. In cavity optomechanics, intracavity photon\nnumber \ructuations disturb the motion of a cantilever,\nproviding the necessary backaction of a quantum mea-\nsurement of position [25]. The analogous disturbance of\noptically probed atomic spins (or pseudo-spins) has been\nstudied both in free-space [26] and intracavity [11, 27]\ncon\fgurations. In an optomechanics-like con\fguration,\ne.g. with B/i, backaction heating of the atomic spin\nenforces quantum limits to measurement of the precess-\ning ensemble and also set limits on optodynamical cool-\ning. In contrast with optomechanical systems, optically\nprobed spin ensembles readily present the opportunity to\nperform quantum-non-demolition (QND) measurements;\nwithB/k, the detected spin component Skis a QND\nvariable representing the energy of the spin system.\nThe noise-perturbed spin acts back upon the cavity\noptical \feld, mediating a self-interaction of the light \feld\n00 . 1-500005000\n246\nt (ms)8.2 8.3 8.4Si and S kSk\nSi\n0\nf (kHz)100 200 300 400-π/201020\n-10π/2(a)\n(b)\nS\nk\nS\ni\n100\n 200\n 300\n 400dBFIG. 3. (Color) Simulations of spin dynamics for S= 5000,\n\nL=2\u0019= 200 kHz, \n c=2\u0019=\u00002:3 kHz,\u0014=2\u0019= 1:8 MHz,\n^n+= 10, and \u0001 p;+= 0:37\u0014. (a) Time evolution of Si(black)\nandSk(blue), following spin preparation near i, shows ampli-\n\fcation, reorientation, and damping toward the high-energy\nstable orientation near \u0000i. Note the di\u000berent scales on the\nhorizontal axis. (b) Logarithmic optical spectral noise power\nrelative to that of coherent light, plotted vs. quadrature an-\ngle\u001e(amplitude quadrature at \u001e= 0), shows inhomogeneous\noptical squeezing. Simulation results shown in color, and lin-\nearized theory (Eq. 10) as contour lines every 5 dB.\nthat can result in optical squeezing. To exhibit this ef-\nfect, we consider a cavity illumined with \u001b+circular po-\nlarized probe light with detuning \u0001 p. The dynamics of\nthe cavity \feld are given by\nd^c+\ndt= (i\u0001p\u0000\u0014+i\nc^Sk)^c++\u0014\u0010\n\u0011+^\u0018+\u0011\n:(8)\nHere,\u0011gives the coherent-state amplitude of the drive\n\feld and the noise operator ^\u0018+represents its \ructua-\ntions. When evaluating the dynamics numerically, we\nconsider a semiclassical Langevin equation, converting\n^\u0018+into a Gaussian stochastic variable with statistics re-\nlated to those of the noise operator, and replacing the\noperators ^c+and^Swithc-numbers. This substitution is\nappropriate for moderately large values of \u0016 nandS.\nFig. 3 portrays the simulated evolution of a spin pre-\npared initially in a low-energy spin orientation (close\ntoi), driven by a blue-detuned cavity probe. Coher-\nent spin ampli\fcation directs the spin toward the stable\nhigh-energy con\fguration (near \u0000i), yielding a dynami-\ncal steady state characterized by a negative temperature.\nTo obtain analytical expressions for the evolution dy-\nnamics, we follow the example of cavity optomechanics [5]4\nby linearizing the Langevin equations for spin and opti-\ncal \ructuations about their steady-state value. The spin\nprojectionSkresponds to amplitude-quadrature \ructua-\ntions of the cavity \feld \u0018A(!) with susceptibility\n\u001f(!)\u0011Sk(!)\n\u0018A(!)=\u0000\nL0\ncp\u0016n+\n\nL02+kSR(!)\u0000!2+i!\u0000o(!);(9)\nwhere \u0000o(!) = 2\u0014\nL02\u0000!2\n\u00142+\u00012\np\u0000!2is the cavity optodynamic\nspin damping, and R(!) =\u00142+\u00012\np\n\u00142+\u00012p\u0000!2. The susceptibil-\nity is largest for !'\nL00. The driven spin feeds the\n\ructuations back onto the cavity \feld, yielding the in-\ntracavity \feld \ructuation spectrum\nc+(!) =\nL02+i\u0014+!\n\u0001pkSR(!)\u0000!2+i!\u0000o(!)\n\nL02+kSR(!)\u0000!2+i!\u0000o(!)\u0018A+i\u0018P;\n(10)\nwhere\u0018P(!) is the input spectrum of phase \ructuations.\nThis \ructuation spectrum exhibits inhomogeneous opti-\ncal squeezing (Fig. 3b).\nApplications: The analogy of cavity optodynamics\nwidens the range of phenomena accessed through the ma-\nnipulation and detection of quantum spins within opti-\ncal cavities, enabling several applications. For example,\nbistability in cavity-coupled single-spin systems serves to\nincrease the readout \fdelity of cavity-coupled qubits [28].\nSimilarly here, cavity-spin bistability could be used as a\nSchmitt trigger for the collective spin: if the probe power\nis turned on diabatically in the bistable regime, the cav-\nity transmission will latch into either a bright or dark\nstate, depending on whether the initial spin state is be-\nlow or above a parametrically chosen threshold value.\nThe cavity-spin system may also be used as a phase-\npreserving ampli\fer for spin dynamics occurring near\nthe shifted precession frequency \n L00, with ampli\fcation\nnoise given in Eq. (10). Both applications may aid mea-\nsurements of ac magnetic \felds, amplifying weak signals\nabove technical sensitivity limits.\nConversely, cavity spin optodynamics may be applied\nas a powerful simulator of cavity optomechanics, with the\nspin system allowing for new means of control. For exam-\nple, precession frequencies may be tuned rapidly by vary-\ning the applied magnetic \feld, simulating optomechanics\nwith a dynamically variable mechanical spring constant.\nAlternately, spatial control of inhomogeneous magnetic\n\felds may be used to divide a spin ensemble into sev-\neral independent subensembles, simulating optomechan-\nics with several mechanical modes.\nIn addition to the dilute gas implementation discussed\nso far, a similar system could be constructed using solid-\nstate spin ensembles and microwave resonators. For ex-\nample, using ensembles of nitrogen-vacancy defects in di-\namond coupled to the circular polarized evanescent radi-\nation of a crossed microwave resonator [29], the Hamilto-\nnian of Eq. (4) is obtained by the ground ms=\u00061 elec-tronic states as the pseudospin and replacing the ac Stark\nshift with an ac Zeeman shift from microwave radiation\nnear the 2.8-GHz crystal-\feld-split Zeeman transition.\nWe thank H. Mabuchi and K.B. Whaley for inspiring\ndiscussions. This work was supported by the NSF and\nthe AFOSR. D.M.S.-K. acknowledges support from the\nMiller Institute for Basic Research in Science.\n\u0003dmsk@berkeley.edu\n[1] T. Kippenberg and K. Vahala, Science 321, 1172 (2008).\n[2] V. Braginskii and F. Y. Khalili, Quantum Measurement\n(Cambridge University Press, Cambridge, 1995).\n[3] K. W. Murch, K. L. Moore, S. Gupta, and D. M.\nStamper-Kurn, Nature Physics 4, 561 (2008).\n[4] V. Vuleti\u0013 c and S. Chu, Phys. Rev. Lett. 84, 3787 (2000).\n[5] C. Fabre et al., Phys. Rev. A 49, 1337 (1994); S. Mancini\nand P. Tombesi, Phys. Rev. A 49, 4055 (1994).\n[6] S. Haroche and J.-M. Raimond, Exploring the Quantum:\nAtoms, Cavities, and Photons (Oxford University Press,\n2006).\n[7] P. F. Herskind et al. , Nature Physics 5, 494 (2009).\n[8] P. E. Barclay, K.-M. C. Fu, C. Santori, and R. G. Beau-\nsoleil, App. Phys. Lett. 95, 191115 (2009).\n[9] L. DiCarlo et al. , Nature 460, 240 (2009).\n[10] J. M. Geremia, J. K. Stockton, A. C. Doherty, and\nH. Mabuchi, Phys. Rev. Lett. 91, 250801 (2003).\n[11] I. D. Leroux, M. H. Schleier-Smith, and V. Vuletic, Phys.\nRev. Lett. 104, 073602 (2010).\n[12] Q. Turchette et al. , Phys. Rev. Lett. 75, 4710 (1995).\n[13] T. Holstein and H. Primako\u000b, Phys. Rev. 58, 1098\n(1940).\n[14] V. B. Braginsky and A. B. Manukin, Sov. Phys. JETP\n25, 653 (1967).\n[15] T.J. Kippenberg et al., Phys. Rev. Lett. 95, 033901\n(2005); S. Gigan et al., Nature 444, 67 (2006); O. Ar-\ncizet et al., Nature 444, 71 (2006).\n[16] A. Dorsel et al. , Phys. Rev. Lett. 51, 1550 (1983).\n[17] S. Gupta, K. L. Moore, K. W. Murch, and D. M.\nStamper-Kurn, Phys. Rev. Lett. 99, 213601 (2007).\n[18] H. J. Kimble et al. , Phys. Rev. D 65, 022002 (2001).\n[19] W. Happer and B. S. Mathur, Phys. Rev. Lett. 18, 577\n(1967).\n[20] C. Cohen-Tannoudji and J. Dupont-Roc, Phys. Rev. A\n5, 968 (1972).\n[21] T. Purdy et al. , preprint, arXiv:1005.4085.\n[22] H. M. Gibbs, Optical bistability: Controlling light with\nlight (Academic Press, New York, 1985).\n[23] A. Buonanno and Y. B. Chen, Phys. Rev. D 65, 042001\n(2002).\n[24] T. Corbitt et al. , Phys. Rev. Lett. 98, 150802 (2007).\n[25] C. M. Caves, Phys. Rev. Lett. 45, 75 (1980).\n[26] C. Schori, B. Julsgaard, J. L. S\u001crensen, and E. S. Polzik,\nPhys. Rev. Lett. 89, 057903 (2002).\n[27] M. H. Schleier-Smith, I. D. Leroux, and V. Vuletic, Phys.\nRev. Lett. 104, 073604 (2010).\n[28] I. Siddiqi et al. , Phys. Rev. B 73, 054510 (2006).\n[29] J. J. Henderson, C. M. Ramsey, H. M. Quddusi, and E. d.\nBarco, Rev. Sci. Instrum. 79, 074704 (2008)." }, { "title": "1608.07646v1.Tunneling_induced_spin_dynamics_in_a_quantum_dot_lead_hybrid_system.pdf", "content": "arXiv:1608.07646v1 [cond-mat.mes-hall] 27 Aug 2016Tunneling induced spindynamics inaquantum dot-leadhybri d system\nTomohiro Otsuka,1,2,∗Takashi Nakajima,1,2Matthieu R. Delbecq,1Shinichi Amaha,1\nJun Yoneda,1,2Kenta Takeda,1Giles Allison,1Peter Stano,1,3Akito Noiri,1,2Takumi\nIto,1,2Daniel Loss,1,4Arne Ludwig,5Andreas D. Wieck,5and Seigo Tarucha1,2,6,7\n1Center for Emergent Matter Science, RIKEN, 2-1 Hirosawa, Wa ko, Saitama 351-0198, Japan\n2Department of Applied Physics, University of Tokyo, Bunkyo , Tokyo 113-8656, Japan\n3Institute of Physics, Slovak Academy of Sciences, 845 11 Bra tislava, Slovakia\n4Department of Physics, University of Basel, Klingelbergst rasse 82, 4056 Basel, Switzerland\n5Angewandte Festk¨ orperphysik, Ruhr-Universit¨ at Bochum , D-44780 Bochum, Germany\n6Quantum-Phase Electronics Center, University of Tokyo, Bu nkyo, Tokyo 113-8656, Japan\n7Institute for Nano Quantum Information Electronics, Unive rsity of Tokyo, 4-6-1 Komaba, Meguro, Tokyo 153-8505, Japan\n(Dated: September 2, 2021)\nSemiconductor quantum dots (QDs)offer a platform toexplor e the physics of quantum electronics including\nspins. Electron spins in QDs are considered good candidates for quantum bits [1] in quantum information\nprocessing [2, 3], and spin control and readout have been est ablished down to a single electron level [4]. We\nuse these techniques to explore spin dynamics in a hybrid sys tem, namely a QD coupled to a two dimensional\nelectronic reservoir. The proximity of the lead results in r elaxation dynamics of both charge and spin, the\nmechanism of which is revealed by comparing the charge and sp in signal. For example, higher order charge\ntunnelingeventscanbemonitoredbyobservingthespin. Wee xpecttheseresultstostimulatefurtherexploration\nof spindynamics inQD-leadhybridsystems and expand the pos sibilities for controlledspin manipulations.\nElectron spins in semiconductor QDs have relatively long\ncoherence times [5–8], while the solid state structures hav e\npotential scalability by utilizing current extensive semi con-\nductor fabricationtechniques. They are also consideredgo od\ncandidates for quantum bits [1] in quantum information pro-\ncessing [2, 3]. The spin states can be initialized using a lar ge\nexchangeenergyofasingleQD[9],andtherelatedPaulispin\nblockadeinadoubleQD(DQD)[10]. Thedemonstratedways\nofmanipulationincludethespin-spinexchangeinteractio nbe-\ntween neighboringQDs [9], and electron spin resonances in-\nducedbymicrocoils[11],nuclearspins[12],spinorbitint er-\naction [13], or micro magnets [14–16]. Finally, the spin can\nbereadout byPaulispin blockade,or a tunnelingsensitive t o\ntheZeemanenergy[17].\nIn experiments aimed at the minimization of the dissipa-\ntion, the QDs have been typically isolated from their envi-\nronment, including the leads, as much as possible. How-\never, it is worth to explore physics in hybrid systems, where\nthe dot-environmentcoupling is stronger, since this coupl ing\ncan be tuned straightforwardly and precisely tuned by gate\nvoltages. In addition, the electronic reservoirs can be tai -\nlored themselves, by applying bias voltages or using specifi c\nstatessuch asferromagnets[18],superconductors[19],qu an-\ntum Hall states [20], and others. This variability gives ris e\nto attractive physics like Fano interference [21, 22], Kond o\nstates[23,24],orgeneralphysicsofopenandnonequilibri um\nsystems,andpossiblyleadtonewwaysofspinmanipulations\nutilizinginteractionsinducedbytheenvironment[25].\nInthiswork,weexplorespindynamicsinaQD-leadhybrid\nsystemutilizingthespinmanipulationandreadouttechniq ues\ndeveloped in previous spin qubit experiments. Specifically ,\nwe monitor changes of the spin and charge states induced by\ncoupling of the QD to an electronic reservoir. With the QD\nbeingclosetoachargetransition,weobservespinandcharg erelaxation,correspondingtofirst-ordertunnelingevents . With\nthedotinaCoulombblockadeconfiguration,weobserveonly\nthe relaxation of spin, corresponding to second-order tunn el-\ningevents.\nFigure 1(a) shows a scanning electron micrograph of the\ndevice. By applyingnegativevoltages on the gate electrode s,\na DQD and a QD charge sensor [26] are formed at the lower\nanduppersides,respectively. TheleftQDintheDQDcouples\nto a lead andthe couplingstrengthis tunedbythe voltage VT\nappliedongateT.TheQDchargesensorisconnectedtoaRF\nresonator formedby the inductor Land the stray capacitance\nCpforRF reflectometry[26–28]. The numberof electronsin\neachQD(n1,n2)ismonitoredbytheintensityofthereflected\nRF signal Vsensor. A change in the electrostatic environment\naround the sensing dot changes its conductance, which shift s\nthe tank circuit resonance and modifies Vsensormeasured at\nfres=297MHz, thecircuitresonancefrequency.\nFigure1(b)showsthechargestabilitydiagramoftheDQD.\nWe measured the sensor signal Vsensoras a function of the\nplunger gate voltages of QD 2(VP2), and QD 1(VP1). We\nobserve a change ∆Vsensoreach time the DQD charge con-\nfiguration (n1,n2)changes. Depicted in Fig. 1(b), the values\n(n1,n2)areassignedbycountingthenumberofchargetransi-\ntionlinesfromthefullydepletedconfiguration (n1,n2)=(0,0)\n[the latter not shown on Fig. 1(b)]. Around the charge state\ntransition (1,1)↔(0,2), we observe a suppression of the\n(0,2)charge signal due to the Pauli spin blockade [in the re-\ngion indicated by the triangle in Fig. 1(b)]. In this specific\nmeasurementofthestability diagram,unlikeelsewhere,up on\npulsing(2,0)→(1,1)we move through the singlet-triplet\nT+anti-crossing very slowly (adiabatically), to induce a siz -\nable triplet component of the (1,1)state even at a zero inter-\naction time. Pulsing quickly back (1,1)→(0,2)results in a\nPauli blocked signal inside the denoted triangular area. Th is2\n(a) (b)\n500 nmP1P2SRF\nL\nCpLead\n-800-790-780-770VP1 (mV)\n-180-170-160-150\nVP2 (mV)-200 200ΔVsensor (mV)\n(1,1) (1,2)\n(0,2) (0,1)SB\nMO1\nIO2\n(c)\nInitialize\nLeadQD1 QD2Operation Measurement\n?\nInteraction SBT z\nyxδ\nFIG. 1: (a) Scanning electron micrograph of the device and th e\nschematic of the measurement setup. A DQD is formed at the low er\nside and the charge states are monitored by the charge sensor QD at\nthe upper side. The charge sensor is connected to resonators formed\nbytheinductor Landthestraycapacitance CpfortheRFreflectom-\netry. The external magnetic field of 0.5 T is applied in plane a long\nthezaxis. (b) ∆Vsensoras a function of VP2andVP1. Changes of\nthe charge states are observed. The number of electrons in ea ch QD\nisgiven as (n1,n2). Thetriangle shows the regionof spinblockade.\nThepositionscorrespondingtostepsofpulsesequences(O, I,M)are\nindicated. (c) Schematic of the measurement scheme. The spi nstate\nis initializedtoa (0,2) singlet at I.Next, the we move intoO in(1,1)\nwherethespincouplestothelead. Finally,thespinstateis measured\nusing spinblockade at M.\nshowsuswherewecanutilizethePaulispinblockadetoread-\nout the spin state in the followingmeasurements,probingth e\ndotspinandchargetunneling-induceddynamics.\nThe operation scheme to measure the effect of the lead on\nthe spin is depicted in Fig. 1(c). We initialize the state to a\n(0,2)singlet by waiting at the initialization point I denot edin\nFig. 1(b). Next, we move to the operation point O. In this\nstep, the electron in QD 1interacts with the lead and the dot\nstate might be changed by electron tunneling. The tunneling\nrate can be modified by tuning VT, which changes the tunnel\ncoupling,andthepositionofO,whichchangesthedotpoten-\ntial with respect to the Fermi energyof the lead (O 1: close to\na charge transition, O 2: deep in the Coulomb blockade). At\nthe next step, the spin state is measured using spin blockade\nby pulsingthe dot to the pointdenotedbyM. If the spin state\ndid notchange,we observethe (0,2)singlet again. If the spi n\nstate changed,a polarized triplet componentis measuredas a\nblocked(1,1)→(0,2)charge transition. From the charge\nsignal,wecanthereforededucethespinstate.\nIn this way, we first measure the spin relaxation using\nthe operation point O 1close to a charge transition line, see\nFig.1(b),wheretheQDlevelisclosetotheFermilevelofthe\nlead. Thetunnelinggatevoltageissetto VT=−660mV.The(b)(a) (c)\n1.0\n0.9\n0.8\n0.7\n0.6\n0.5Singlet Probability\n12840\nTime (µs)-920-900-880-860\nCharge Signal (mV)\n12\n8\n4\n0Time (µs)\n-1200 -800 -400\nVsensor (mV)12x10-3\n8\n4\n0(1,1) (0,1)Lead QD1\nEmptyN(Vsensor)/Ntot\nFIG.2: (a) Observed spin and charge signals (the singlet pro bability\nand the average of the sensor signal /angbracketleftVsensor/angbracketright) as a function of the\ninteraction time. Red circles show the spin signal (left axi s). The\nblue trace shows the charge signal (right axis). The smooth l ines are\nexponential fitsresultingintherelaxationtimeof3.0 µsforthespin,\nand 1.8µs for the charge. (b) Statistics of the charge signal at the\noperation point. Histogram of observed values of the charge sensor\nVsensor(on the x axis), N(Vsensor)/Ntotis plotted as a function of\nthe interactiontime(yaxis). The twopeaks, at Vsensor=−960mV,\nand−780mV, correspond to the (1,1) and the (0,1) charge states,\nrespectively. The weight of the (0,1) component increases w ith the\nlonger interaction time. (c) Schematic of the spin relaxati on by a\nfirst-order tunneling process. An electron escapes from the QD and\nthe QDbecomes empty. Another electroncomes inafterthat.\nred circles in Fig. 2(a) show the measured singlet probabili ty\nas a function of the interaction time at O 1. Initially at 1, the\nsinglet probability decreases upon increasing the interac tion\ntime from zero. This decrease indicates that a triplet compo -\nnentisformedbytheinteractionwiththelead. Fittingwith an\nexponentialrevealsa relaxationtimeof3.0 µs.\nSimilarly to spin, we also measure the lifetime of charge\nin this configuration. The blue trace in Fig. 2(a) shows the\naveraged Vsensorasafunctionoftheinteractiontime. Asseen\nthere,∝an}bracketle{tVsensor∝an}bracketri}htchanges exponentially, with the fitted charge\nrelaxationtimeof1.8 µs. Toexaminethechargerelaxationin\nmore detail, we plot in Fig. 2b histograms (the x axis) of the\nvalues of Vsensorfor a varying interaction time (the y axis).\nThe two peaks along a horizontal cut correspond to the (1,1)\nand the (0,1)chargestates, respectively. At a zero interac tion\ntime, only the (1,1) state signal is present, while (0,1) sta te\nappearsforfiniteinteractiontimes.\nIn this configuration, the mechanism of the relaxation for\nboth spin and charge is a first-order tunneling process [29].3\n12\n8\n4\n0Time (µs)\n-1000 -800 -600\nVsensor (mV)20x10-3\n15\n10\n5\n00.9\n0.8\n0.7Singlet Probability\n12840\nTime (µs)-800-780-760-740-720\nCharge Signal (mV)\n(b)(a) (c)\n(1,1)Lead QD1N(Vsensor)/Ntot\nFIG.3: (a) The observed singlet probability and /angbracketleftVsensor/angbracketrightas a func-\ntion of the interaction time at O 2[see Fig. 1(b)]. Red circles show\nthe spin signal (left axis). The blue trace shows the charge s ignal\n(right axis). The red smooth curve is an exponential fit resul ting in\ntherelaxationtimeof4.5 µs. Thecharge signalshowsnorelaxation.\n(b) Histogram of observed values of the charge sensor Vsensor(on\nthe x axis), N(Vsensor)/Ntot, is plotted as a function of the interac-\ntion time (y axis). The peak corresponds to the (1,1) charge s tate.\n(c)Schematicofthespinrelaxationbyasecond-order tunne lingpro-\ncess. An electron of the QD 1is swapped with a one in the lead in a\nsingle step. The spinstateis changed even though the charge state is\nstable.\nNamely, the electron tunnels out of the QD 1into the lead,\nafter which the dot is refilled from the lead, and the initial\ninformationislost. Thespinandchargerelaxationhappens i-\nmultaneously,theinformationlossofthespindemonstrate din\nFig.2(a),andofthechargeinFig.2(a-b). Wenotethatthoug h\nthe relaxation timescales are similar, they are not identic al.\nThedifferencecomesfromadifferenceintheratedependenc e\nonFermioccupationofthelead(seetheSupplementaryInfor -\nmation).\nWenowinvestigatethespindynamicsinaCoulombblock-\nadeddot. Tothisend,werepeatthepreviouslydescribedmea -\nsurement using the operation point O 2, deep in the (1,1) re-\ngion, see Fig. 1(b). Here the QD level is far below the Fermi\nlevel of the lead. To increase the speed of the lead induced\nspindynamicsonthedot,weincreasethedot-leadtunnelcou -\npling by setting VT=−560mV. As can be seen in Fig. 3(a),\nsimilarly as before, the spin state displays an exponential de-\ncay, with the relaxation time of 4.5 µs. However, now the\ncharge signal barely changes, indicating that the charge st ate\nis not affected. (The slight change of the charge signal in\nFig. 3(a) is caused by the distorted voltage pulses applied o n2.5\n2.0\n1.5\n1.0\n0.5\n0.0Relaxation Rate (MHz)\n543210\nδ (mV)\n0.9\n0.8\n0.7Singlet Probability\n543210\nTime (µs)(a)\n(b)\nFIG. 4: (a) The spin relaxation rate as a function of δ. Circles show\nthe experimental data and the line shows a theoretical curve consid-\neringthesecond-ordertunnelingprocess. (b)Thespinrela xationsig-\nnal as a function of the interaction time at O2for different values of\nthe gate voltage applied on gate T VT. Circles,triangles and squares\nshow the result at VT=−560,−565,−570mV, respectively. The\nlines are exponential fits.\nP1 and P2. Due to a cross-talk between the plunger gates\nandthesensor,thepulsedistortionslightlyaffecttheobs erved\nchargesignal.) This is confirmedby Fig. 3(b), where the his-\ntograms of the values of Vsensordisplay a single peak corre-\nsponding to the (1,1) charge state. The spin therefore decay s\nata fixedQDchargeconfiguration.\nWe therefore interpret this as the observation of a spin re-\nlaxation induced by a second-order tunneling process [30],\nwhere the electron in QD 1swaps with a random one from\ntheleadinasinglestep. Figure4(a)showsthespinrelaxati on\nrate as we change the operation point from O 2toward O 1,\nparametrizingthedisplacementbyvoltage δ. Uponincreasing\nδ(moving towards the charge transition line) the spin relax-\nationrateisenhanced. Themeasureddependenceisverywell\nfitted by an analytical expression for an inelastic cotunnel ing\nrate,giving ∝(1/(µ(2)−µF)+1/(µF−µ(1)))2,withµ(N)\nandµFbeing the electrochemical potential at the dot with N\nelectrons [4] and the Fermi energy of the lead, respectively\n(see the SupplementalInformationfor details). In additio n to\nplungervoltages,wecantunethe spindecaytimescalebythe\nvoltage applied on gate T, VT. Indeed, as seen in Fig. 4, ap-\nplyingmore negativevoltage VTprolongsthe spin relaxation\ntime,bydecreasingthetunnelcouplingtothelead,from0.7 to\n1.7 to 5.0 µs, forVT=−560,−565,−570mV, respectively.\n(We notethatthe relaxationtime at VT=−560mV isdiffer-4\nentfromthecorrespondingvalueof VTgiveninFig.3(a)due\nto a shift of the QD conditions between experiments.) This\ndemonstratesthetwohandlesonthespeedoftheleadinduced\ndynamicsoftheQD spin.\nTo sum up the results observed in the Coulomb blockade\nregime, here the interaction with the lead influences only th e\ndot spin, but not charge. The spin relaxationthusdirectly u n-\ncoversthe second ordertunnelingprocesses. This interact ion\ncanbeutilizedforthespininitialization,thoughalsothe mea-\nsurement and manipulation might be envisioned, considerin g\nleads with special properties. We note that even though the\ntimescale of the dot-lead interaction realized in this expe ri-\nment was tuned to ∼µs, it is straightforward to enhance it\nby increasing the tunnel coupling, and/or utilizing the Kon do\neffect,whichenhancesthesecond-ordertunnelingatlowte m-\nperatures.\nIn conclusion, we have measured spin dynamics in a QD-\nleadhybridsystem. Close toa chargetransition,we observe d\nspin and charge relaxation signals corresponding to the firs t-\norder tunneling process. In the Coulomb blockade, we ob-\nserved spin relaxation at a fixed charge configuration, corre -\nsponding to the second-ordertunneling process. The demon-\nstrated dot-lead spin exchange can be useful as a general re-\nsource for spin manipulations, and simulations of open sys-\ntemsundernon-equilibriumconditions.\nMETHODS\nThe device was fabricated from a GaAs/AlGaAs het-\nerostructure wafer with an electron sheet carrier density o f\n2.0×1015m−2and a mobility of 110 m2/Vs at 4.2 K, mea-\nsured by Hall-effect in the van der Pauw geometry. The two-\ndimensional electron gas is formed 90 nm under the wafer\nsurface. We patterned a mesa by wet-etching and formed\nTi/Au Schottky surface gates by metal deposition, which ap-\npear white in Fig. 1(a). All measurementswere conductedin\na dilutionfridgecryostatat a temperatureof13mK.\nACKNOWLEDGEMENTS\nWe thank J. Beil, J. Medford, F. Kuemmeth, C. M. Mar-\ncus, D. J. Reilly, K. Ono, RIKEN CEMS Emergent Mat-\nter Science ResearchSupportTeamandMicrowaveResearch\nGroup in Caltech for fruitful discussions and technical sup -\nports. Part of this work is supported by the Grant-in-Aid for\nScientific Research (No. 25800173, 26220710, 26709023,\n26630151, 16H00817), CREST, JST, ImPACT Program of\nCouncilforScience,TechnologyandInnovation(CabinetOf -\nfice, Government of Japan), Strategic Information and Com-\nmunications R&D Promotion Programme, RIKEN Incentive\nResearch Project, Yazaki Memorial Foundation for Science\nand TechnologyResearch Grant, Japan Prize FoundationRe-\nsearchGrant,AdvancedTechnologyInstituteResearchGran t,the Murata Science Foundation Research Grant, Izumi Sci-\nence and Technology Foundation Research Grant, TEPCO\nMemorialFoundationResearchGrant,IARPAproject“Multi-\nQubit Coherent Operations” throughCopenhagenUniversity ,\nDFG-TRR160,andtheBMBF - Q.com-H16KIS0109.\nAUTHOR CONTRIBUTIONS\nT. O., T. N., M. D., S. A., J. Y., K. T., G. A. and S. T.\nplannedtheproject;T.O.,T.N.,M.D.,S.A.,A.L.andA.W.\nperformeddevicefabrication;T. O., T.N., M. D., S. A., J. Y. ,\nK. T., G. A., P. S., A. N., T. I., D. L. and S. T. conductedex-\nperimentsand data analysis; all authorsdiscussed the resu lts;\nT. O., T. N., M. D., S. A., J. Y., K. T., G. A., P. S. and S. T.\nwrotethemanuscript.\n∗tomohiro.otsuka@riken.jp\n[1] Loss, D., and DiVincenzo, D. P. Quantum computation with\nquantum dots. Phys.Rev. A 57, 120(1998).\n[2] Nielsen, M. A., and Chuang, I. L. Quantum Computation and\nQuantum Information. (Cambridge UniversityPress,2000).\n[3] Ladd, T. D. et al. Quantum computers. Nature 464, 45-53\n(2010).\n[4] Hanson, R. et al. Spins in few-electron quantum dots. Rev.\nMod. Phys. 79, 1217 (2007).\n[5] Bluhm, H. et al. Dephasing time of GaAs electron-spin qubits\ncoupled to a nuclear bath exceeding 200 µs. Nat. Phys. 7, 109-\n113 (2011).\n[6] Shulman, M. D. et al. Suppressing qubit dephasing using real-\ntime Hamiltonianestimation. Nat.Commun. 5,5156 (2014).\n[7] Kawakami, E. etal.Electricalcontrolofalong-livedspinqubit\nina Si/SiGequantum dot.Nat. Nano. 9, 666-670 (2014).\n[8] Veldhorst, M. et al. An addressable quantum dot qubit with\nfault-tolerant control-fidelity.Nat. Nano. 9, 981-985 (2014).\n[9] Petta, J. R. et al. Coherent Manipulation of Coupled Electron\nSpins in Semiconductor Quantum Dots. Science 309, 2180-\n2184 (2005).\n[10] Ono, K., Austing, D. G., Tokura, Y., and Tarucha, S. Curr ent\nRectification by Pauli Exclusion in a Weakly Coupled Double\nQuantum Dot System.Science 297, 1313-1317 (2002).\n[11] Koppens, F.H.L. etal. Drivencoherent oscillations of a single\nelectron spinina quantum dot. Nature 442, 766-771 (2006).\n[12] Laird, E. A. et al. Hyperfine-Mediated Gate-Driven Electron\nSpinResonance. Phys. Rev. Lett. 99, 246601 (2007).\n[13] Nowack, K. C., Koppens, F. H. L., Nazarov, Y. V., and Van-\ndersypen, L.M. K. Coherent Control of a Single Electron Spin\nwithElectricFields.Science 318, 1430-1433 (2007).\n[14] Tokura, Y., van der Wiel, W. G., Obata, T., and Tarucha, S .\nCoherent Single Electron Spin Control in a Slanting Zeeman\nField.Phys. Rev. Lett. 96, 047202 (2006).\n[15] Pioro-Ladriere,M. etal.Electricallydrivensingle-electronspin\nresonance in a slanting Zeeman field. Nat. Phys. 4, 776-779\n(2008).\n[16] Yoneda,J. etal.FastElectricalControlofSingleElectronSpins\ninQuantumDotswithVanishingInfluencefromNuclearSpins.\nPhys. Rev. Lett. 113, 267601 (2014).5\n[17] Elzerman,J.M. etal.Single-shotread-outofanindividualelec-\ntronspinina quantum dot. Nature 430, 431-435 (2004).\n[18] Hamaya, K. et al. Spin transport through a single self-\nassembled InAs quantum dot with ferromagnetic leads. Appl.\nPhys.Lett. 90, 053108 (2007).\n[19] Deacon,R.S. etal.TunnelingSpectroscopyofAndreevEnergy\nLevels in a Quantum Dot Coupled to a Superconductor. Phys.\nRev. Lett. 104, 076805 (2010).\n[20] Altimiras,C. etal.Non-equilibriumedge-channelspectroscopy\nintheintegerquantum Hallregime.Nat.Phys. 6,34-39(2010).\n[21] Fano,U.Effectsof ConfigurationInteractiononIntens itiesand\nPhase Shifts.Phys. Rev. 124, 1866-1878 (1961).\n[22] Kobayashi, K., Aikawa, H., Katsumoto, S., and Iye, Y. Tu ning\nof the Fano Effect through a Quantum Dot in an Aharonov-\nBohm Interferometer. Phys.Rev. Lett. 88, 256806 (2002).\n[23] Goldhaber-Gordon, D. et al. Kondo effect in a single-electron\ntransistor.Nature 391, 156-159 (1998).\n[24] van der Wiel, W. G. et al. The Kondo Effect in the Unitary\nLimit.Science 289, 2105-2108 (2000).\n[25] Anders, F. B., and Schiller, A. Spin precession and real -time\ndynamics in the Kondo model: Time-dependent numerical\nrenormalization-group study. Phys.Rev. B 74, 245113 (2006).\n[26] Barthel,C. etal.Fastsensingofdouble-dotchargearrangement\nandspinstatewitharadio-frequencysensorquantumdot.Ph ys.\nRev. B81, 161308 (2010).\n[27] Schoelkopf, R. J. et al. The Radio-Frequency Single-Electron\nTransistor (RF-SET): A Fast and Ultrasensitive Electromet er.\nScience280, 1238-1242 (1998).\n[28] Reilly, D. J., Marcus, C. M., Hanson, M. P., and Gossard, A.\nC. Fast single-charge sensing with a rf quantum point contac t.\nAppl.Phys. Lett. 91, 162101 (2007).\n[29] Biesinger, D. E. F. et al. Intrinsic Metastabilities in the Charge\nConfiguration of a Double Quantum Dot Phys. Rev. Lett. 115,\n106804 (2015).\n[30] Schleser,R. etal.Cotunneling-MediatedTransportthroughEx-\ncitedStatesintheCoulomb-Blockade Regime.Phys.Rev.Let t.\n94, 206805 (2005).1\nSupplementalMaterialto`Tunneling induced spin dynamics in a quantumdot-leadhybridsystem'\nCHARGEAND SPINRELAXATIONSIGNALS\nWe describethedynamicsofthechargeandspinontheQD bycon sideringtherateequation\n∂tPσ=−Γσ(1−fσ)Pσ+ΓσfσPe, (S1)\nfor the probability Pσthat the dot is occupied by a single electron with spin σ∈ {↑,↓}(we alternative use σ=±1), where\nPeis the probability that the dot is empty. We do not consider an y other states, which gives the normalization condition\nP↑+P↓+Pe= 1. Eq. (S1) includesthe processof an electron with a given spi n leaving the dot into the lead where an empty\nstateexistswiththeprobability (1−fσ)andenteringanemptydotfromtheleadstateoccupiedwithpr obability fσ. ThisFermi\nfactor is given by fσ=fFD(µ(1)−σEz), withµ(1) =eVg+ǫ1being the energycost to add an electron into the dot, which\nincludesthe electrostatic potentialenergy eVg, the orbital (quantization)energy ǫ1. The Zeemanenergyis Ez=gµBB/2, and\ntheFermi-Diracdistribution\nfFD(ǫ) =/braceleftbigg\nexp/bracketleftbiggǫ−µF\nkBT/bracketrightbigg\n+1/bracerightbigg−1\n, (S2)\ndependson the temperature T, and the lead Fermi energy µF. Apart fromthe Fermi factors fσ, the tunnelingrates for hopping\nonandoffthe dotare identical, Γσ. We, however,allowfora spin dependenceofthe tunnelingra tewhichhasbeenfoundto be\nanappreciableeffect(theasymmetryoftheratescanbeofth eorderoftheratesthemselves),mostprobablyduetotheexc hange\ninteractioninthelead[1–3].\nTo expose the spin and charge dynamics, we introduce new vari ables, the probability of charge occupation, Po=P↑+P↓\nand the spin polarization, s=P↑−P↓, and new parameters, for the average, Γ, and the dimensionless asymmetry α, in the\ntunnelingrates, bywriting Γσ= Γ(1+ σα), andsimilarly fortheFermi factors, fσ=f+σfδ. Equation(S1) canbe nowcast\nintothematrixformforthe vectorofunknowns, v= (Po,s)T,namely\n∂tv=−M(v−v0), (S3)\nwiththe matrixdefiningthesystem propagator\nM= Γ/parenleftbigg\n1+f+fδα−fδ+(1−f)α\nfδ+(1+f)α1−f−fδα/parenrightbigg\n, (S4)\nandthesteadystate solution\nv0=2\n1−f2+f2\nδ/parenleftbigg\nf(1−f)−f2\nδ\nfδ/parenrightbigg\n. (S5)\nThesteadystateisindependentonthetunnelingrates,andd ependsonlyontheleadFermifactorsforthetwospins,asit s hould\nbe,whilethepropagatormatrixdependsonallparametersof theproblem. Eventhoughitisstraightforwardtosolvethep roblem\ninthe mostgeneralcase,it isusefultoconsider Mforα= 0(spinindependenttunnelingrates),whichgives\nM= Γ/parenleftbigg\n1+f−fδ\nfδ1−f/parenrightbigg\n. (S6)\nForanegligibledifferenceoftheFermifunctionvaluesfor thetwospins, fδ→0,thechargeandspindecaytotheirsteadystate\nvalues independently,with the rates Γ(1+f), andΓ(1−f), respectively. The steady states are also markedly differe ntin this\nlimit, asPo(t=∞) = 2f/(1+f)depends on the Fermi factors, while s(t=∞) = 0does not. This is then the reason for\ndifferenceinthedecayscales: whilethechargedecaystowa rdsthesteadystate witheffectivelythesumoftheratesfor leaving,\n(1−f)Γ,andentering, 2fΓ,thedot,onlytheeventsofelectronsleavingthedotscanre laxthespinpolarization s. Thespinand\nchargerelaxationscaleswill thenbemostdifferentif f≈1,wherethechargeequilibratesmuchfaster thanthespin.\nTodemonstratethisdifference,seenalsoexperimentally, weplotthechargeandspinsignalsinFig.S1(a)and(b),resp ectively.\nTheparametersare set as T= 200mK,B= 0.5T,g=−0.37[4]. Thetracesshowtheresultswith f= 0.2,0.4,0.6,0.8from\nthebottomto thetop.2\n1.0\n0.8\n0.6\n0.4Charge Signal\n543210(a)\nt(1/Γ)1.0\n0.9\n0.8\n0.7\n0.6Spin Signal\n543210(b)\nt(1/Γ)\nFIG.S1: (a) Calculatedcharge signal asa functionof t. Thetraces show the resultswith f= 0.2,0.4,0.6,0.8from thebottom tothe top. (b)\nCalculatedspinsignal as afunction of t.\nCO-TUNNELINGRATE\nTo derive the formula for the spin relaxation by cotunneling , which was used in the main text to fit the data on Fig. 4(a), we\nconsidertheHamiltonianofa QDcoupledtoa lead, H=HD+HL+HT. HerethedotHamiltonianis\nHD=/summationdisplay\nα∈{0,σ,S}ǫα|α∝an}bracketri}ht∝an}bracketle{tα|, (S7)\nwheretheindex αlabelsthestatesofthedot |α∝an}bracketri}htwithenergies ǫα,and|0∝an}bracketri}htdenotesanemptydot, |σ∝an}bracketri}ht=d†\nσ|0∝an}bracketri}htadotwithasingle\nelectronwith spin σ, and|S∝an}bracketri}ht=d†\n↑d†\n↓|0∝an}bracketri}hta dot with a two electronsinglet state, and d†\nσis the creation operatorof a dot electron\nwithspin σ. Theleadisdescribedby\nHL=/summationdisplay\nkσǫkσc†\nkσckσ, (S8)\nwherekisawave-vector(forsimplicity,weconsideraonedimensio nallead,sothat kisascalar). Finally,thedot-leadcoupling\nis\nHT=/summationdisplay\nkσtkσc†\nkσdσ+t∗\nkσd†\nσckσ, (S9)\nwhich desctribes a spin-preserving lead-dot tunneling wit h, in general complex and spin and energy dependent, tunneli ng am-\nplitudestkσ.\nWe now repeat the standard calculation [5–9] with minor adju stments to to arrive at the inelastic spin decay rather than t he\nco-tunnelingcurrent. Tothisend,wedefinethetransitionr ate bythe Fermi’sGoldenruleformula\nΓi→f=/summationdisplay\nileadpilead/summationdisplay\nflead2π\n/planckover2pi1|∝an}bracketle{tfdot⊗flead|G|idot⊗ilead∝an}bracketri}ht|2δ(Ei−Ef), (S10)\nwhereiandfaretheinitialandfinalstateswithenergies EiandEf,respectively,consideredtobeseparable(totheleadandd ot\ncomponents) eigenstates of the unperturbed system describ ed byH0=HD+HL. As we are not conditioning the transitions\non the states of the lead, the rate is summed over all possible initial lead states, with the correspondingprobabilities pilead, and\nall lead final states. The former gives the prescription for a replacement/summationtext\nileadpilead|ilead∝an}bracketri}ht∝an}bracketle{tilead| →ρthermal\nlead, with the latter\nthe equilibriumdensity matrixcorrespondingto a system wi th Hamiltonian HL, at a temperature T. Finally, Gis the transition\noperatorwhichcanbeexpandedinpowersofthetunnelingter m\nG=HT+HT1\nE−H0−iγHT+... (S11)\nwithE=Ei=Ef. The two terms describe, respectively, the direct tunnelin g and the co-tunneling, and γis a regularization\nfactor[10].3\nSimple results can be derived in the well justified case of a ne gligible dependence of the tunneling amplitudes on the wave\nvector,tkσ≈tσ. Using the first term in Eq. (S11) gives in this limit the follo wing expression for the direct tunneling rates\ndefinedin Eq.(S1)\nΓσ=2π\n/planckover2pi1|tσ|2gF, (S12)\nwhere we denoted Γσ≡Γσ→0= Γ0→σandgFis the density of states in the lead at the Fermi energy. Simil arly, keeping\nonlythe secondterminEq.(S11)givestheco-tunnelingrate fora spin-flip(from σto theoppositevalue σ)ofa singleelectron\noccupyingthedot,\nΓσ→σ=/planckover2pi1\n2πΓσΓσ/integraldisplay\ndǫfFD(ǫ+σEz)[1−fFD(ǫ+σEz)]/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\nǫ−µ(2)+iγ−1\nǫ−µ(1)−iγ/vextendsingle/vextendsingle/vextendsingle/vextendsingle2\n, (S13)\nwhereµ(2) =ǫS−ǫ1isthe(spinindependentpartofthe)energycosttoaddaseco ndelectronintothedot. Theexpressioncan\nbe furthersimplified if the dot is deep in the Coulombblockad e,so that the chargeexcitation energiesare muchlarger tha n the\ntemperature,namely µ(1)≪µF≪µ(2)arewellfulfilledontheenergyscaleofthetemperature, kBT. Theenergydependence\nofthelast terminEq.(S13)canbe thenneglected,replacing ǫ→µF, andtheremainingintegralcanbeevaluatedresultingin\nΓσ→σ=/planckover2pi1\n2πΓσΓσ2σEz\nexp/parenleftBig\n2σEz\nkBT−1/parenrightBig/parenleftbigg1\nµ(2)−µF+1\nµF−µ(1)/parenrightbigg2\n, (S14)\nwherewealsoneglectedtheregularizationfactors. Inthel argetemperaturelimit, kBT≫Ez,thetemperaturedependentfactor\nbecomeskBT,whileintheoppositelimit, kBT≫Ez,itgives2Ezforσ=↓,and0forσ=↑. However,Eq.(S14)isalreadyin\ntheformwhichwasusedtofit the dataandisthusthefinalresul tofthissection.\n∗tomohiro.otsuka@riken.jp\n[S1] S. Amasha, K. MacLean, Iuliana P. Radu, D. M. Zumb¨ ul, M. A. Kastner, M. P. Hanson, and A. C. Gossard, Spin-dependent tunneling\nof single electrons intoan empty quantum dot , Phys.Rev. B 78, 041306(R) (2008).\n[S2] P.StanoandPh. Jacquod, Spin-dependent tunneling intoan empty lateral quantum dot , Phys.Rev. B 82, 125309 (2010).\n[S3] M.Yamagishi,N.Watase,M.Hashisaka,K.Muraki,andT. Fujisawa, Spin-dependent tunnelingratesforelectrostaticallydefi nedGaAs\nquantum dots , Phys. Rev. B 90, 035306 (2014).\n[S4] T. Otsuka, T. Nakajima, M. R. Delbecq, S. Amaha, J. Yoned a, K. Takeda, G. Allison, T. Ito, R. Sugawara, A. Noiri, A. Lud wig, A. D.\nWieckand S.Tarucha, Single-electronSpin Resonance inaQuadruple Quantum Dot , Sci.Rep. 6, 31820 (2016).\n[S5] D. V. Averin and Yu. V. Nazarov, Virtual electron diffusion during quantum tunneling of the electric charge , Phys. Rev. Lett. 65, 2446\n(1990).\n[S6] M. R.Wegewijs,and Yu.V. Nazarov, Inelastic co-tunneling through anexcitedstate of aquantu m dot, arXiv:cond-mat/0103579.\n[S7] V. N. Golovach and D. Loss, Transport through a double quantum dot in the sequential tun neling and cotunneling regimes , Phys. Rev.\nB69, 245327 (2004).\n[S8] J.Lehmann andD.Loss, Cotunneling current throughquantum dots withphonon-assi sted spin-flipprocesses , Phys.Rev. B73, 045328\n(2006)\n[S9] P. Stano, J. Klinovaja, F. R. Braakman, L. M. K. Vandersy pen, and D. Loss, Fast Long-Distance Control of Spin Qubits by Photon\nAssistedCotunneling , Phys.Rev. B 92, 075302 (2015).\n[S10] G. Begemann, S.Koller, M. Grifoni,andJ. Paaske, Inelastic cotunneling inquantum dots and molecules withwe akly broken degenera-\ncies, Phys.Rev. B 82, 045316 (2010)." }, { "title": "1012.3575v1.Spin_Relaxation_in_Quantum_Wires.pdf", "content": "arXiv:1012.3575v1 [cond-mat.mes-hall] 16 Dec 2010Spin Relaxation in Quantum Wires\nP. Wenk∗\nSchool of Engineering and Science, Jacobs University Breme n, Bremen 28759, Germany\nS. Kettemann†\nSchool of Engineering and Science, Jacobs University Breme n, Bremen 28759,\nGermany, and Asia Pacific Center for Theoretical Physics and Division of Advanced\nMaterials Science Pohang University of Science and Technol ogy (POSTECH) San31,\nHyoja-dong, Nam-gu, Pohang 790-784, South Korea\nThe spin dynamics and spin relaxation of itinerant electron s in quantum wires with spin-orbit\ncoupling is reviewed. We give an introduction to spin dynami cs, and review spin-orbit coupling\nmechanisms in semiconductors. The spin diffusion equation w ith spin-orbit coupling is derived,\nusing only intuitive, classical random walk arguments. We g ive an overview of all spin relaxation\nmechanisms, with particular emphasis on the motional narro wing mechanism in disordered conduc-\ntors, the D’yakonov-Perel’-Spin relaxation (DPS). Here, w e discuss in particular, the existence of\npersistent spin helix solutions of the spin diffusion equati on, with vanishing spin relaxation rates.\nWe then, derive solutions of the spin diffusion equation in qu antum wires, and show that there is an\neffective alignment of the spin-orbit field in wires whose wid th is smaller than the spin precession\nlengthLSO. We show that the resulting reduction in the spin relaxation rate results in a change\nin the sign of the quantum corrections to the conductivity. F inally, we present recent experimental\nresults which confirm the decrease of the spin relaxation rat e in wires whose width is smaller than\nLSO: the direct optical measurement of the spin relaxation rate , as well as transport measurements,\nwhich show a dimensional crossover from weak antilocalizat ion to weak localization as the wire\nwidth is reduced. Open problems remain, in particular in nar rower, ballistic wires, were optical and\ntransport measurements seem to find opposite behavior of the spin relaxation rate: enhancement,\nsuppression, respectively. We conclude with a review of the se and other open problems which still\nchallenge the theoretical understanding and modeling of th e experimental results.2\nContents\nI. Introduction 3\nII. Spin Dynamics 3\nA. Dynamics of a Localized Spin 3\nB. Spin Dynamics of Itinerant Electrons 4\n1. Ballistic Spin Dynamics 4\n2. Spin Diffusion Equation 4\n3. Spin-Orbit Interaction in Semiconductors 5\n4. Spin Diffusion in the Presence of Spin-Orbit Interaction 7\nIII. Spin Relaxation Mechanisms 9\nA. D’yakonov-Perel’ Spin Relaxation 10\nB. DP Spin Relaxation with Electron-Electron and Electron- Phonon Scattering 11\nC. Elliott-Yafet Spin Relaxation 12\nD. Spin Relaxation due to Spin-Orbit Interaction with Impur ities 12\nE. Bir-Aronov-Pikus Spin Relaxation 13\nF. Magnetic Impurities 13\nG. Nuclear Spins 13\nH. Magnetic Field Dependence of Spin Relaxation 13\nI. Dimensional Reduction of Spin Relaxation 14\nIV. Spin-Dynamics in Quantum Wires 14\nA. One-Dimensional Wires 14\nB. Spin-Diffusion in Quantum Wires 14\nC. Weak Localization Corrections 17\nV. Experimental Results on Spin Relaxation Rate in Semicond uctor Quantum Wires 19\nA. Optical Measurements 19\nB. Transport Measurements 20\nVI. Critical Discussion and Future Perspective 20\nVII. Summary 21\nSymbols 22\nAcknowledgments 22\nReferences 223\nI. INTRODUCTION\nThe emerging technology of spintronics intends to use the ma nipulation of the spin degree of freedom of individual\nelectrons for energy efficient storage and transport of infor mation.1In contrast to classical electronics, which relies on\nthe steering of charge carriers through semiconductors, sp intronics uses the spin carried by electrons, resembling ti ny\nspinning tops. The difference to a classical top is that its an gular momentum is quantized, it can only take two discrete\nvalues, up or down. To control the spin of electrons, a detail ed understanding of the interaction between the spin and\norbital degrees of freedom of electrons and other mechanism s which do not conserve its spin, is necessary. These are\ntypically weak perturbations, compared to the kinetic ener gy of conduction electrons, so that their spin relaxes slowl y\nto the advantage of spintronic applications. The relaxatio n, or depolarization of the electron spin can occur due to\nthe randomization of the electron momentum by scattering fr om impurities, and dislocations in the material, and due\nto scattering with elementary excitations of the solid such as phonons and other electrons, when it is transferred to\nthe randomization of the electron spin due to the spin-orbit interaction. In addition, scattering from localized spins ,\nsuch as nuclear spins and magnetic impurities are sources of electron spin relaxation. The electron spin relaxation can\nbe reduced by constraining the electrons in low dimensional structures, quantum wells (confined in one direction, free\nin two dimensions), quantum wires ( confined in two direction s, free in one direction), or quantum dots ( confined in\nall three directions). Although spin relaxation is typical ly smallest in quantum dots due to their discrete energy leve l\nspectrum, the necessity to transfer the spin in spintronic d evices, recently lead to intense research efforts to reduce t he\nspin relaxation in quantum wires, where the energy spectrum is continuous. In the following we will review the theory\nof spin dynamics and relaxation in quantum wires, and compar e it with recent experimental results. After a general\nintroduction to spin dynamics in Section II, we discuss all relevant spin relaxation mechanisms and how they depend\non dimension, temperature, mobility, charge carrier densi ty and magnetic field in Section III. In particular, we review\nrecent results on spin relaxation in semiconducting quantu m wires, and its influence on the quantum corrections to\ntheir conductance in Section IV. These weak localization corrections are thereby a very sen sitive measure of spin\nrelaxation in quantum wires, in addition to optical methods as we review in Section V. We set /planckover2pi1= 1in the following.\nII. SPIN DYNAMICS\nBefore we review the spin dynamics of conduction electrons a nd holes in semiconductors and metals, let us first\nreconsider the spin dynamics of a localized spin, as governe d by the Bloch equations.\nA. Dynamics of a Localized Spin\nA localized spin ˆ s, like a nuclear spin, or the spin of a magnetic impurity in a so lid, precesses in an external\nmagnetic field Bdue to the Zeeman interaction with Hamiltonian HZ=−γgˆ sB, whereγgis the corresponding\ngyromagnetic ratio of the nuclear spin or magnetic impurity spin, respectively, which we will set equal to one, unless\nneeded explicitly. This spin dynamics is governed by the Blo ch equation of a localized spin,\n∂tˆ s=γgˆ s×B. (1)\nThis equation is identical to the Heisenberg equation ∂tˆ s=−i[ˆ s,HZ]for the quantum mechanical spin operator ˆ sof\nanS= 1/2-spin, interacting with the external magnetic field Bdue to the Zeeman interaction with Hamiltonian HZ.\nThe solution of the Bloch equation for a magnetic field pointi ng in the z-direction is ˆsz(t) = ˆsz(0), while the x- and y-\ncomponents of the spin are precessing with frequency ω0=γgBaround the z-axis, ˆsx(t) = ˆsx(0)cosω0t+ˆsy(0)sinω0t,\nˆsy(t) =−ˆsx(0)sinω0t+ ˆsy(0)cosω0t. Since a localized spin interacts with its environment by ex change interaction\nand magnetic dipole interaction, the precession will depha se after a time τ2, and the z-component of the spin relaxes\nto its equilibrium value sz0within a relaxation time τ1. This modifies the Bloch equations to the phenomenological\nequations,\n∂tˆsx=γg(ˆsyBz−ˆszBy)−1\nτ2ˆsx\n∂tˆsy=γg(ˆszBx−ˆsxBz)−1\nτ2ˆsy\n∂tˆsz=γg(ˆsxBy−ˆsyBx)−1\nτ1(ˆsz−sz0). (2)4\nB. Spin Dynamics of Itinerant Electrons\n1. Ballistic Spin Dynamics\nThe intrinsic degree of freedom spin is a direct consequence of the Lorentz invariant formulation of quantum\nmechanics. Expanding the relativistic Dirac equation in th e ratio of the electron velocity and the constant velocity\nof lightc, one obtains in addition to the Zeeman term, a term which coup les the spin swith the momentum pof the\nelectrons, the spin-orbit coupling\nHSO=−µB\n2mc2ˆ s p×E=−ˆ sBSO(p), (3)\nwhere we set the gyromagnetic ratio γg= 1.E=−∇V, is an electrical field, and BSO(p) =µB/(2mc2)p×E.\nSubstitution into the Heisenberg equation yields the Bloch equation in the presence of spin-orbit interaction:\n∂tˆ s=ˆ s×BSO(p), (4)\nso that the spin performs a precession around the momentum de pendent spin-orbit field BSO(p). It is important to\nnote, that the spin-orbit field does not break the invariance under time reversal ( ˆ s→ −ˆ s,p→ −p), in contrast to\nan external magnetic field B. Therefore, averaging over all directions of momentum, the re is no spin polarization of\nthe conduction electrons. However, injecting a spin-polar ized electron with given momentum pinto a translationally\ninvariant wire, its spin precesses in the spin-orbit field as the electron moves through the wire. The spin will be\noriented again in the initial direction after it moved a leng thLSO, the spin precession length. The precise magnitude\nofLSOdoes not only depend on the strength of the spin-orbit intera ction but may also depend on the direction of its\nmovement in the crystal, as we will discuss below.\n2. Spin Diffusion Equation\nTranslational invariance is broken by the presence of disor der due to impurities and lattice imperfections in the\nconductor. As the electrons scatter from the disorder poten tial elastically, their momentum changes in a stochastic\nway, resulting in diffusive motion. That results in a change o f the the local electron density ρ(r,t) =/summationtext\nα=±|ψα(r,t)|2,\nwhereα=±denotes the orientation of the electron spin, and ψα(r,t)is the position and time dependent electron\nwave function amplitude. On length scales exceeding the ela stic mean free path le, that density is governed by the\ndiffusion equation\n∂ρ\n∂t=De∇2ρ, (5)\nwhere the diffusion constant Deis related to the elastic scattering time τbyDe=v2\nFτ/dD, wherevFis the Fermi\nvelocity, and dDthe Diffusion dimension of the electron system. That diffusio n constant is related to the mobility\nof the electrons, µe=eτ/m∗by the Einstein relation µeρ=e2νDe, whereνis the density of states per spin\nat the Fermi energy EF. Injecting an electron at position r0into a conductor with previously constant electron\ndensityρ0, the solution of the diffusion equation yields that the elect ron density spreads in space according to\nρ(r,t) =ρ0+exp(−(r−r0)2/4Det)/(4πDet)dD/2, wheredDis the dimension of diffusion. That dimension is equal to\nthe kinetic dimension d,dD=d, if the elastic mean free path leis smaller than the size of the sample in all directions.\nIf the elastic mean free path is larger than the sample size in one direction the diffusion dimension reduces by one,\naccordingly. Thus, on average the variance of the distance t he electron moves after time tis/angb∇acketleft(r−r0)2/angb∇acket∇ight= 2dDDet. This\nintroduces a new length scale, the diffusion length LD(t) =√Det. We can rewrite the density as ρ=/angb∇acketleftψ†(r,t)ψ(r,t)/angb∇acket∇ight,\nwhereψ†= (ψ†\n+,ψ†\n−)is the two-component vector of the up (+), and down (-) spin fe rmionic creation operators, and\nψthe 2-component vector of annihilation operators, respect ively,/angb∇acketleft.../angb∇acket∇ightdenotes the expectation value. Accordingly,\nthe spin density s(r,t)is expected to satisfy a diffusion equation, as well. The spin density is defined by\ns(r,t) =1\n2/angb∇acketleftψ†(r,t)σψ(r,t)/angb∇acket∇ight, (6)\nwhereσis the vector of Pauli matrices,\nσx=/parenleftbigg\n0 1\n1 0/parenrightbigg\n,σy=/parenleftbigg\n0−i\ni0/parenrightbigg\n,andσz=/parenleftbigg\n1 0\n0−1/parenrightbigg\n.5\nThus the z-component of the spin density is half the differenc e between the density of spin up and down electrons,\nsz= (ρ+−ρ−)/2, which is the local spin polarization of the electron system . Thus, we can directly infer the diffusion\nequation for sz, and, similarly, for the other components of the spin densit y, yielding, without magnetic field and\nspin-orbit interaction,2\n∂s\n∂t=De∇2s−s\nˆτs. (7)\nHere, in the spin relaxation term we introduced the tensor ˆτs, which can have non-diagonal matrix elements. In the\ncase of a diagonal matrix, τsxx=τsyy=τ2, is the spin dephasing time, and τszz=τ1the spin relaxation time.\nThe spin diffusion equation can be written as a continuity equ ation for the spin density vector, by defining the spin\ndiffusion current of the spin components si,\nJsi=−De∇si. (8)\nThus, we get the continuity equation for the spin density com ponentssi,\n∂si\n∂t+∇Jsi=−/summationdisplay\njsj\nτsij. (9)\n3. Spin-Orbit Interaction in Semiconductors\nWhile silicon and germanium have in their diamond structure an inversion symmetry around every midpoint on each\nline connecting nearest neighbor atoms, this is not the case for III-V-semiconductors like GaAs, InAs, InSb, or ZnS.\nThese have a zinc-blende structure which can be obtained fro m a diamond structure with neighbored sites occupied by\nthe two different elements. Therefore the inversion symmetr y is broken, which results in spin-orbit coupling. Similarl y,\nthat symmetry is broken in II-VI-semiconductors. This bulk inversion asymmetry (BIA) coupling, or often so called\nDresselhaus-coupling, is anisotropic, as given by3\nHD=γD/bracketleftbig\nσxkx(k2\ny−k2\nz)+σyky(k2\nz−k2\nx)+σzkz(k2\nx−k2\ny)/bracketrightbig\n, (10)\nwhereγDis the Dresselhaus-spin-orbit coefficient. Confinement in qu antum wells with width aon the order of the\nFermi wave length λFyields accordingly a spin-orbit interaction where the mome ntum in growth direction is of the\norder of 1/a. Because of the anisotropy of the Dresselhaus term, the spin -orbit interaction depends strongly on the\ngrowth direction of the quantum well. Grown in [001]direction, one gets, taking the expectation value of Eq. ( 10) in\nthe direction normal to the plane, noting that /angb∇acketleftkz/angb∇acket∇ight=/angb∇acketleftk3\nz/angb∇acket∇ight= 0,3\nHD[001]=α1(−σxkx+σyky)+γD(σxkxk2\ny−σykyk2\nx). (11)\nwhereα1=γD/angb∇acketleftk2\nz/angb∇acket∇ightis the linear Dresselhaus parameter. Thus, inserting an ele ctron with momentum along the x-\ndirection, with its spin initially polarized in z-directio n, it will precess around the x-axis as it moves along. For nar row\nquantum wells, where /angb∇acketleftk2\nz/angb∇acket∇ight ∼1/a2≥k2\nFthe linear term exceeds the cubic Dresselhaus terms. A speci al situation\narises for quantum wells grown in the [110]- direction, where it turns out that the spin-orbit field is po inting normal\nto the quantum well, as shown in Fig. 1, so that an electron whose spin is initially polarized along the normal of the\nplane, remains polarized as it moves in the quantum well. In q uantum wells with asymmetric electrical confinement\nthe inversion symmetry is broken as well. This structural in version asymmetry (SIA) can be deliberately modified\nby changing the confinement potential by application of a gat e voltage. The resulting spin-orbit coupling, the SIA\ncoupling, also called Rashba-spin-orbit interaction4is given by\nHR=α2(σxky−σykx), (12)\nwhereα2depends on the asymmetry of the confinement potential V(z)in the direction z, the growth direction of the\nquantum well, and can thus be deliberately changed by applic ation of a gate potential. At first sight it looks as if the\nexpectation value of the electrical field Ec=−∂zV(z)in the conduction band state vanishes, since the ground stat e of\nthe quantum well must be symmetric in z. Taking into account the coupling to the valence band,5,6the discontinuities\nin the effective mass,7and corrections due to the coupling to odd excited states,8yields a sizable coupling parameter\ndepending on the asymmetry of the confinement potential6,9.\nThis dependence allows one, in principle, to control the ele ctron spin with a gate potential, which can therefore be\nused as the basis of a spin transistor.106\nSIA\n00 BIA/LBracket1111/RBracket1\n00\nBIA/LBracket1001/RBracket1\n00\n/LBracket1100/RBracket1/LBracket1010/RBracket1\nBIA/LBracket1110/RBracket1\n0\n/LBracket11/OverBar/OverBar10/RBracket10/LBracket1001/RBracket10/LBracket1110/RBracket1\nFigure 1: The spin-orbit vector fields for linear structure i nversion asymmetry (Rashba) coupling, and for linear bulk i nversion\nasymmetry (BIA) spin orbit coupling for quantum wells grown in [111], [001] and [110] direction, respectively.\nWe can combine all spin-orbit couplings by introducing the s pin-orbit field such that the Hamiltonian has the form\nof a Zeeman term:\nHSO=−sBSO(k), (13)\nwhere the spin vector is s=σ/2. But we stress again that since BSO(k)→BSO(−k) =−BSO(k)under the time\nreversal operation, spin-orbit coupling does not break tim e reversal symmetry, since the time reversal operation also\nchanges the sign of the spin, s→ −s. Only an external magnetic field Bbreaks the time reversal symmetry. Thus,\nthe electron spin operator ˆ sis for fixed electron momentum kgoverned by the Bloch equations with the spin-orbit\nfield,\n∂ˆ s\n∂t=ˆ s×(B+BSO(k))−1\nˆτsˆ s. (14)\nThe spin relaxation tensor is no longer necessarily diagona l in the presence of spin-orbit interaction.\nIn narrow quantum wells where the cubic Dresselhaus couplin g is weak compared to the linear Dresselhaus and Rashba\ncouplings, the spin-orbit field is given by\nBSO(k) =−2\n−α1kx+α2ky\nα1ky−α2kx\n0\n, (15)\nwhich changes both its direction and its amplitude |BSO(k)|= 2/radicalbig\n(α2\n1+α2\n2)k2−4α1α2kxky, as the direction of\nthe momentum kis changed. Accordingly, the electron energy dispersion cl ose to the Fermi energy is in general7\nanisotropic as given by\nE±=1\n2m∗k2±αk/radicalbigg\n1−4α1α2\nα2cosθsinθ, (16)\nwherek=|k|,α=/radicalbig\nα2\n1+α2\n2, andkx=kcosθ. Thus, when an electron is injected with energy E, with momentum\nkalong the [100]-direction, kx=k,ky= 0, its wave function is a superposition of plain waves with the positive\nmomentak±=∓αm∗+m∗(α2+2E/m∗)1/2. The momentum difference k−−k+= 2m∗αcauses a rotation of the\nelectron eigenstate in the spin subspace. When at x= 0the electron spin was polarized up spin, with the Eigenvecto r\nψ(x= 0) =/parenleftbigg\n1\n0/parenrightbigg\n,\nthen, when its momentum points in x-direction, at a distance x, it will have rotated the spin as described by the\nEigenvector\nψ(x) =1\n2/parenleftbigg1\nα1+iα2\nα/parenrightbigg\neik+x+1\n2/parenleftbigg1\n−α1+iα2\nα/parenrightbigg\neik−x. (17)\nIn Fig. 2we plot the corresponding spin density as defined in Eq. ( 6) for pure Rashba coupling, α1= 0. The spin\n0 LSO/Slash12 LSO/Minus101\nxsz\nFigure 2: Precession of a spin injected at x= 0, polarized in z-direction, as it moves by one spin precessio n length LSO=π/m∗α\nthrough the wire with linear Rashba spin orbit coupling α2.\nwill point again in the initial direction, when the phase diff erence between the two plain waves is 2π, which gives the\ncondition for spin precession length as 2π= (k−−k+)LSO, yielding for linear Rashba and Dresselhaus coupling, and\nthe electron moving in [100]- direction,\nLSO=π/m∗α. (18)\nWe note that the period of spin precession changes with the di rection of the electron momentum since the spin-orbit\nfield, Eq. ( 15), is anisotropic.\n4. Spin Diffusion in the Presence of Spin-Orbit Interaction\nAs the electrons are scattered by imperfections like impuri ties and dislocations, their momentum is changed ran-\ndomly. Accordingly, the direction of the spin-orbit field BSO(k)changes randomly as the electron moves through the\nsample. This has two consequences: the electron spin direct ion becomes randomized, dephasing the spin precession\nand relaxing the spin polarization. In addition, the spin pr ecession term is modified, as the momentum kchanges\nrandomly, and has no longer the form given in the ballistic Bl och-like equation, Eq. ( 14). One can derive the diffu-\nsion equation for the expectation value of the spin, the spin density Eq. ( 6) semiclassically,11,12or by diagrammatic8\nexpansion.13In order to get a better understanding on the meaning of this e quation, we will give a simplified classical\nderivation, in the following. The spin density at time t+∆tcan be related to the one at the earlier time t. Note that\nfor ballistic times ∆t≤τ, the distance the electron has moved with a probability p∆x,∆x, is related to that time\nby the ballistic equation, ∆x=k(t)∆t/mwhen the electron moves with the momentum k(t). On this time scale the\nspin evolution is still governed by the ballistic Bloch equa tion Eq. ( 14). Thus, we can relate the spin density at the\nposition xat the time t+∆t, to the one at the earlier time tat position x−∆x:\ns(x,t+∆t) =/summationdisplay\n∆xp∆x/parenleftbigg/parenleftbigg\n1−1\nˆτs∆t/parenrightbigg\ns(x−∆x,t)−∆t[B+BSO(k(t))]×s(x−∆x,t)/parenrightbigg\n. (19)\nNow, we can expand in ∆tto first order and in ∆xto second order. Next, we average over the disorder potentia l,\nassuming that the electrons are scattered isotropically, a nd substitute/summationtext\n∆xp∆x...=/integraltext\n(dΩ/Ω)...whereΩis the total\nangle, and/integraltextdΩdenotes the integral over all angles with/integraltext(dΩ/Ω) = 1 . Also, we get (s(x,t+∆t)−s(x,t))/∆t→\n∂ts(x,t)for∆t→0, and/angb∇acketleft∆x2\ni/angb∇acket∇ight= 2De∆t, whereDeis the diffusion constant. While the disorder average yields\n/angb∇acketleft∆x/angb∇acket∇ight= 0, and/angb∇acketleftBSO(k(t))/angb∇acket∇ight= 0, separately, for isotropic impurity scattering, averagin g their product yields a finite\nvalue, since ∆xdepends on the momentum at time t,k(t), yielding /angb∇acketleft∆xBSOi(k(t))/angb∇acket∇ight= 2∆t/angb∇acketleftvFBSOi(k(t))/angb∇acket∇ight, where\n/angb∇acketleft.../angb∇acket∇ightdenotes the average over the Fermi surface. This way, we can a lso evaluate the average of the spin-orbit term in\nEq. (19), expanded to first order in ∆x, and get, substituting ∆t→τthe spin diffusion equation,\n∂s\n∂t=−B×s+De∇2s+2τ/angb∇acketleft(∇vF)BSO(p)/angb∇acket∇ight×s−1\nˆτss, (20)\nwhere/angb∇acketleft.../angb∇acket∇ightdenotes the average over the Fermi surface. Spin polarized e lectrons injected into the sample spread\ndiffusively, and their spin polarization, while spreading d iffusively as well, decays in amplitude exponentially in tim e.\nSince, between scattering events the spins precess around t he spin-orbit fields, one expects also an oscillation of the\npolarization amplitude in space. One can find the spatial dis tribution of the spin density which is the solution of\nEq. (20) with the smallest decay rate Γs. As an example, the solution for linear Rashba coupling is,12\ns(x,t) = (ˆeqcosqx+Aˆezsinqx)e−t/τs, (21)\nwith1/τs= 7/16τs0where1/τs0= 2τk2\nFα2\n2and where the amplitude of the momentum qis determined by Deq2=\n15/16τs0, andA= 3/√\n15, andˆeq=q/q. This solution is plotted in Fig. 3forˆeq= (1,1,0)/√\n2. In Fig. 4we plot the\nlinearly independent solution obtained by interchanging cosandsinin Eq. ( 21), with the spin pointing in z-direction,\ninitially. We choose ˆeq= ˆex. Comparison with the ballistic precession of the spin, Fig 4shows that the period\nof precession is enhanced by the factor 4/√\n15in the diffusive wire, and that the amplitude of the spin densi ty is\nmodulated, changing from 1toA= 3/√\n15.\n0\n/LParen115/Slash14/RParen1LSO/Slash12\n/LParen115/Slash14/RParen1LSOx/Minus0.500.5Sy\n/Minus0.500.5\nSz\nFigure 3: The spin density for linear Rashba coupling which i s a solution of the spin diffusion equation with the relaxatio n\nrate7/16τs. The spin points initially in the x−y-plane in the direction (1,1,0).9\n0 Lso/Slash12 Lso/Minus/FractionBarExt1\n20/FractionBarExt1\n2\nx/LessS/Greaterz\n0.770.891/VertBar1S/VertBar1\nFigure 4: The spin density for linear Rashba coupling which i s a solution of the spin diffusion equation with the relaxatio n rate\n1/τs= 7/16τs0. Note that, compared to the ballistic spin density, Fig. 2, the period is slightly enhanced by a factor 4/√\n15.\nAlso, the amplitude of the spin density changes with the posi tionx, in contrast to the ballistic case. The color is changing in\nproportion to the spin density amplitude.\nInjecting a spin-polarized electron at one point, say x= 0, its density spreads the same way it does without spin-\norbit interaction, ρ(r,t) = exp(−r2/4Det)/(4πDet)dD/2, whereris the distance to the injection point. However, the\ndecay of the spin density is periodically modulated as a func tion of2π/radicalbig\n15/16r/LSO.14The spin-orbit interaction\ntogether with the scattering from impurities is also a sourc e of spin relaxation, as we discuss in the next Section\ntogether with other mechanisms of spin relaxation. We can fin d the classical spin diffusion current in the presence of\nspin-orbit interaction, in a similar way as one can derive th e classical diffusion current: The current at the position\nris a sum over all currents in its vicinity which are directed t owards that position. Thus, j(r,t) =/angb∇acketleftvρ(r−∆x)/angb∇acket∇ight\nwhere an angular average over all possible directions of the velocity vis taken. Expanding in ∆x=lev/v, and\nnoting that /angb∇acketleftvρ(r)/angb∇acket∇ight= 0, one gets j(r,t) =/angb∇acketleftv(−∆x)∇ρ(r)/angb∇acket∇ight=−(vFle/2)∇ρ(r) =−De∇ρ(r). For the classical spin\ndiffusion current of spin component Si, as defined by jSi(r,t) =vSi(r,t), there is the complication that the spin\nkeeps precessing as it moves from r−∆xtor, and that the spin-orbit field changes its direction with the direction\nof the electron velocity v. Therefore, the 0-th order term in the expansion in ∆xdoes not vanish, rather, we get\njSi(r,t) =/angb∇acketleftvSk\ni(r,t)/angb∇acket∇ight −De∇Si(r,t), whereSk\niis the part of the spin density which evolved from the spin den sity\natr−∆xmoving with velocity vand momentum k. Noting that the spin precession on ballistic scales t≤τis\ngoverned by the Bloch equation, Eq. ( 14), we find by integration of Eq. ( 14), thatSk\ni=−τ(BSO(k)×S)iso that we\ncan rewrite the first term yielding the total spin diffusion cu rrent as\njSi=−τ/angb∇acketleftvF(BSO(k)×S)i/angb∇acket∇ight−De∇Si. (22)\nThus, we can rewrite the spin diffusion equation in terms of th is spin diffusion current and get the continuity equation\n∂si\n∂t=−De∇jSi+τ/angb∇acketleft∇vF(BSO(k)×S)i/angb∇acket∇ight−1\nˆτsijsj. (23)\nIt is important to note that in contrast to the continuity equ ation for the density, there are two additional terms, due\nto the spin orbit interaction. The last one is the spin relaxa tion tensor which will be considered in detail in the next\nsection. The other term arises due to the fact that Eq. ( 20) contains a factor 2in front of the spin-orbit precession\nterm, while the spin diffusion current Eq. ( 22) does not contain that factor. This has important physical c onsequences,\nresulting in the suppression of the spin relaxation rate in q uantum wires and quantum dots as soon as their lateral\nextension is smaller than the spin precession length LSO, as we will see in the subsequent Sections.\nIII. SPIN RELAXATION MECHANISMS\nThe intrinsic spin-orbit interaction itself causes the spi n of the electrons to precess coherently, as the electrons\nmove through a conductor, defining the spin precession lengt hLSO, Eq. ( 18). Since impurities and dislocations in\nthe conductor randomize the electron momentum, the impurit y scattering is transferred into a randomization of the10\nelectron spin by the spin-orbit interaction, which thereby results in spin dephasing and spin relaxation. This results\nin a new length scale, the spin relaxation length, Ls, which is related to the spin relaxation rate 1/τsby\nLs=/radicalbig\nDeτs. (24)\nA. D’yakonov-Perel’ Spin Relaxation\nD’yakonov-Perel’ spin relaxation (DPS) can be understood q ualitatively in the following way: The spin-orbit field\nBSO(k)changes its direction randomly after each elastic scatteri ng event from an impurity, that is, after a time of\nthe order of the elastic scattering time τ, when the momentum is changed randomly as sketched in Fig. 5. Thus, the\nFigure 5: Elastic scattering from impurities changes the di rection of the spin-orbit field around which the electron spi n is\nprecessing.\nspin has the time τto perform a precession around the present direction of the s pin-orbit field, and can thus change\nits direction only by an angle of the order of BSOτby precession. After a time twithNt=t/τscattering events,\nthe direction of the spin will therefore have changed by an an gle of the order of |BSO|τ√Nt=|BSO|√\nτt. Defining\nthe spin relaxation time τsas the time by which the spin direction has changed by an angle of order one, we thus\nfind that 1/τs∼τ/angb∇acketleftBSO(k)2/angb∇acket∇ight, where the angular brackets denote integration over all ang les. Remarkably, this spin\nrelaxation rate becomes smaller, the more scattering event s take place, because the smaller the elastic scattering tim e\nτis, the less time the spin has to change its direction by prece ssion. Such a behavior is also well known as motional ,\nordynamic narrowing of magnetic resonance lines.15A more rigorous derivation for the kinetic equation of the sp in\ndensity matrix yields additional interference terms, not t aken into account in the above argument. It can be obtained\nby iterating the expansion of the spin density Eq. ( 19) once in the spin precession term, which yields the term\n/angbracketleftBigg\ns(x,t)×/integraldisplay∆t\n0dt′BSO(k(t′))×/integraldisplay∆t\n0dt′′BSO(k(t′′))/angbracketrightBigg\n, (25)\nwhere/angb∇acketleft.../angb∇acket∇ightdenotes the average over all angles due to the scattering fro m impurities. Since the electrons move\nballistically at times smaller than the elastic scattering time, the momenta are correlated only on time scales smaller\nthanτ, yielding /angb∇acketleftki(t′)kj(t′′)/angb∇acket∇ight= (1/2)k2δijτδ(t′−t′′).\nNoting that (A×B×C)m=ǫijkǫklmAiBjCland/summationtextǫijkǫklm=δilδjm−δimδjlwe find that Eq. ( 25) simplifies to\n−/summationtext\ni(1/τsij)Sj, where the matrix elements of the spin relaxation terms are g iven by,16\n1\nτsij=τ/parenleftbig\n/angb∇acketleftBSO(k)2/angb∇acket∇ightδij−/angb∇acketleftBSO(k)iBSO(k)j/angb∇acket∇ight/parenrightbig\n, (26)\nwhere/angb∇acketleft.../angb∇acket∇ightdenotes the average over the direction of the momentum k. These non-diagonal terms can diminish the\nspin relaxation and even result in vanishing spin relaxatio n. As an example, we consider a quantum well where the\nlinear Dresselhaus coupling for quantum wells grown in [001]direction, Eq. ( 11), and linear Rashba-coupling, Eq. ( 12),\nare the dominant spin-orbit couplings. The energy dispersi on is anisotropic, as given by Eq. ( 16), and the spin-orbit\nfieldBSO(k)changes its direction and its amplitude with the direction o f the momentum k:\nBSO(k) =−2\n−α1kx+α2ky\nα1ky−α2kx\n0\n, (27)11\nwith|BSO(k)|= 2/radicalbig\n(α2\n1+α2\n2)k2−4α1α2kxky. Thus we find the spin relaxation tensor as,\n1\nˆτs(k) = 4τk2\n1\n2α2−α1α20\n−α1α21\n2α20\n0 0 α2\n. (28)\nDiagonalizing this matrix, one finds the three eigenvalues (1/τs)(α1±α2)2/α2and2/τswhereα2=α2\n1+α2\n2, and\n1/τs= 2k2τα2. Note, that one of these eigenvalues of the spin relaxation t ensor vanishes when α1=α2=α0. In\nfact, this is a special case, when the spin-orbit field does no t change its direction with the momentum:\nBSO(k)|α1=α2=α0=2α0(kx−ky)\n1\n1\n0\n. (29)\nIn this case the constant spin density given by\nS=S0\n1\n1\n0\n, (30)\ndoes not decay in time, since the spin density vector is paral lel to the spin orbit field BSO(k), Eq. ( 29), and cannot\nprecess, as has been noted in Ref. [ 17]. It turns out, however, that there are two more modes which d o not decay in\ntime, whose spin relaxation rate vanishes for α1=α2. These modes are not homogeneous in space, and correspond to\nprecessing spin densities. They were found previously in a n umerical Monte Carlo simulation and found not to decay in\ntime, being called therefore persistent spin helix .18,19Recently, a long living inhomogeneous spin density distrib ution\nhas been detected experimentally in Ref. [ 20]. We can now get these persistent spin helix modes analytically, by solving\nthe full spin diffusion equation Eq. ( 20) with the spin relaxation tensor given by Eq. ( 28). We can diagonalize that\nequation, noting that its eigenfunctions are plain waves S(x)∼exp(iQx−Et). Thereby one finds, first of all, the\nmode with Eigenvalue E1=DeQ2, with the spin density\nS=S0\n1\n1\n0\nexp(iQx−DeQ2t). (31)\nIndeed for Q= 0, the homogeneous solution, it does not decay in time, in agre ement with the solution we found\nabove, Eq. ( 31). There are, however, two more modes with the eigenvalues\nE±=1\nτs(˜Q2+2±2|˜Qx−˜Qy|), (32)\nwhere˜Q=LSOQ/2π. At˜Qx=−˜Qy=±1, these modes do not decay in time. These two stationary solut ions, are\nS=S0\n1\n−1\n0\nsin/parenleftbigg2π\nLSO(x−y)/parenrightbigg\n+S0√\n2\n0\n0\n1\ncos/parenleftbigg2π\nLSO(x−y)/parenrightbigg\n, (33)\nand the linearly independent solution, obtained by interch angingcosandsinin Eq. ( 33). The spin precesses as the\nelectrons diffuse along the quantum wire with the period LSO, the spin precession length, forming a persistent spin\nhelix, as shown in Fig. 6.\nB. DP Spin Relaxation with Electron-Electron and Electron- Phonon Scattering\nIt has been noted, that the momentum scattering which limits the D’yakonov-Perel’ mechanism of spin relaxation\nis not restricted to impurity scattering, but can also be due to electron-phonon or electron-electron interactions.21–24\nThus the scattering time, τis the total scattering time as defined by,21,221/τ= 1/τ0+1/τee+1/τep, where1/τ0is the\nelastic scattering rate due to scattering from impurities w ith potential V, given by 1/τ0= 2πνni/integraltext\n(dθ/2π)(1−cosθ)|\nV(k,k′)|2, whereνis the density of states per spin at the Fermi energy, niis the concentration of impurities with\npotentialV, andkk′=kk′cos(θ). In degenerate semiconductors and in metals, the electron- electron scattering rate\nis given by the Fermi liquid inelastic electron scattering r ate1/τee∼T2/ǫF. The electron-phonon scattering time\n1/τep∼T5decays faster with temperature. Thus, at low temperatures t he DP spin relaxation is dominated by elastic\nimpurity scattering τ0.12\n0\nLSO/Slash12\nLSOx\n/MinusS00S0\n/LessS/Greatery/Minus/Radical1/Radical1Extens/Radical1Extens/Radical1Extens/Radical1Extens/Radical1Extens2S00/Radical1/Radical1Extens/Radical1Extens/Radical1Extens/Radical1Extens/Radical1Extens2S0\n/LessS/Greaterz\nFigure 6: Persistent spin helix solution of the spin diffusio n equation for equal magnitude of linear Rashba and linear Dr esselhaus\ncoupling, Eq.( 33).\nC. Elliott-Yafet Spin Relaxation\nBecause of the spin-orbit interaction the conduction elect ron wave functions are not Eigenstates of the electron\nspin, but have an admixture of both spin up and spin down wave f unctions. Thus, a nonmagnetic impurity potential\nVcan change the electron spin, by changing their momentum due to the spin-orbit coupling. This results in another\nsource of spin relaxation which is stronger, the more often t he electrons are scattered, and is thus proportional to the\nmomentum scattering rate 1/τ.25,26For degenerate III-V semiconductors one finds27,28\n1\nτs∼∆2\nSO\n(EG+∆SO)2E2\nk\nE2\nG1\nτ(k), (34)\nwhereEGis the gap between the valence and the conduction band of the s emiconductor, Ekthe energy of the\nconduction electron, and ∆SOis the spin-orbit splitting of the valence band. Thus, the El liott-Yafet spin relaxation\n(EYS) can be distinguished, being proportional to 1/τ, and thereby to the resistivity, in contrast to the DP spin\nscattering rate, Eq. ( 26), which is proportional to the conductivity. Since the EYS d ecays in proportion to the inverse\nof the band gap, it is negligible in large band gap semiconduc tors likeSiandGaAs . The scattering rate 1/τis again\nthe sum of the impurity scattering rate,25the electron-phonon scattering rate,26,29and electron-electron interaction,30\nso that all these scattering processes result in EYS. In non- degenerate semiconductors, where the Fermi energy is\nbelow the conduction band edge, 1/τs∼τT3/EGattains a stronger temperature dependence.\nD. Spin Relaxation due to Spin-Orbit Interaction with Impur ities\nThe spin-orbit interaction, as defined in Eq. ( 3), arises whenever there is a gradient in an electrostatic po tential.\nThus, the impurity potential gives rise to the spin-orbit in teraction\nVSO=1\n2m2c2∇V×k s. (35)\nPerturbation theory yields then directly the correspondin g spin relaxation rate\n1\nτs=πνni/summationdisplay\nα,β/integraldisplaydθ\n2π(1−cosθ)|VSO(k,k′)αβ|2, (36)\nproportional to the concentration of impurities ni. Hereα,β=±denotes the spin indices. Since the spin-orbit\ninteraction increases with the atomic number Zof the impurity element, this spin relaxation increases as Z2, being\nstronger for heavier element impurities.13\nE. Bir-Aronov-Pikus Spin Relaxation\nThe exchange interaction Jbetween electrons and holes in p-doped semiconductors resu lts in spin relaxation, as\nwell.31,32Its strength is proportional to the density of holes pand depends on their itinerancy. If the holes are\nlocalized they act like magnetic impurities. If they are iti nerant, the spin of the conduction electrons is transferred by\nthe exchange interaction to the holes, where the spin-orbit splitting of the valence bands results in fast spin relaxati on\nof the hole spin due to the Elliott-Yafet, or the D’yakonov-P erel’ mechanism.\nF. Magnetic Impurities\nMagnetic impurities have a spin Swhich interacts with the spin of the conduction electrons by the exchange\ninteraction J, resulting in a spatially and temporarily fluctuating local magnetic field\nBMI(r) =−/summationdisplay\niJδ(r−Ri)S, (37)\nwhere the sum is over the position of the magnetic impurities Ri. This gives rise to spin relaxation of the conduction\nelectrons, with a rate given by\n1\nτMs= 2πnMνJ2S(S+1), (38)\nwherenMis the density of magnetic impurities, and νis the density of states at the Fermi energy. Here, Sis the\nspin quantum number of the magnetic impurity, which can take the valuesS= 1/2,1,3/2,2.... Antiferromagnetic\nexchange interaction between the magnetic impurity spin an d the conduction electrons results in a competition\nbetween the conduction electrons to form a singlet with the i mpurity spin, which results in enhanced nonmagnetic\nand magnetic scattering. At low temperatures the magnetic i mpurity spin is screened by the conduction electrons\nresulting in a vanishing of the magnetic scattering rate. Th us, the spin scattering from magnetic impurities has a\nmaximum at a temperature of the order of the Kondo temperatur eTK∼EFexp(−1/νJ), whereνis the density of\nstates at the Fermi energy.33–35In semiconductors TKis exponentially small due to the small effective mass and the\nresulting small density of states ν. Therefore, the magnetic moments remain free at the experim entally achievable\ntemperatures. At large concentration of magnetic impuriti es, the RKKY-exchange interaction between the magnetic\nimpurities quenches however the spin quantum dynamics, so t hatS(S+1)is replaced by its classical value S2. In\nMn-p-doped GaAs, the exchange interaction between the Mn do pants and the holes can result in compensation of the\nhole spins and therefore a suppression of the Bir-Aronov-Pi kus (BAP) spin relaxation.36\nG. Nuclear Spins\nNuclear spins interact by the hyperfine interaction with con duction electrons. The hyperfine interaction between\nnuclear spins ˆIand the conduction electron spin, ˆs, results in a local Zeeman field given by37\nˆBN(r) =−8π\n3g0µB\nγg/summationdisplay\nnγnˆI,δ(r−Rn), (39)\nwhereγnis the gyromagnetic ratio of the nuclear spin. The spatial an d temporal fluctuations of this hyperfine\ninteraction field result in spin relaxation proportional to its variance, similar to the spin relaxation by magnetic\nimpurities.\nH. Magnetic Field Dependence of Spin Relaxation\nThe magnetic field changes the electron momentum due to the Lo rentz force, resulting in a continuous change of\nthe spin-orbit field, which similar to the momentum scatteri ng results in motional narrowing and thereby a reduction\nof DPS:28,38–40\n1\nτs∼τ\n1+ω2cτ2. (40)14\nAnother source of a magnetic field dependence is the precessi on around the external magnetic field. In bulk semi-\nconductors and for magnetic fields perpendicular to a quantu m well, the orbital mechanism is dominating, however.\nThis magnetic field dependence can be used to identify the spi n relaxation mechanism, since the EYS does have only\na weak magnetic field dependence due to the weak Pauli-parama gnetism.\nI. Dimensional Reduction of Spin Relaxation\nElectrostatic confinement of conduction electrons can redu ce the effective dimension of their motion. In quantum\ndots, the electrons are confined in all three directions, and the e nergy spectrum consists of discrete levels like in\natoms. Therefore, the energy conservation restricts relax ation processes severely, resulting in strongly enhanced s pin\nrelaxation times in quantum dots.41,42Then, spin relaxation can only occur due to absorption or emi ssion of phonons,\nyielding spin relaxation rates proportional to the inelast ic electron-phonon scattering rate.41Quantitative comparison\nof the various spin relaxation mechanisms in GaAs quantum do ts resulted in the conclusion that the spin relaxation is\ndominated by the hyperfine interaction.43–45A similar conclusion can be drawn from experiments on low tem perature\nspin relaxation in low density n-type GaAs, where the locali zation of the electrons in the impurity band results in\nspin relaxation dominated by hyperfine interaction as well.46,47For linear Rashba and linear Dresselhaus spin-orbit\ncoupling we can see from the spin diffusion equation Eq. ( 20) with the DP spin relaxation tensor Eq. ( 28) that the\nspin relaxation vanishes, when the spin current Eq. ( 22) vanishes, in which case the last two terms of Eq. ( 20)\ncancel exactly. The vanishing of the spin current is imposed by hard wall boundary condition for which the spin\ndiffusion current vanishes at the boundaries of the sample, jSin|Boundary= 0, wherenis the normal to the boundary.\nWhen the quantum dot is smaller than the spin precession leng thLSOthe lowest energy mode thus corresponds to a\nhomogeneous solution with vanishing spin relaxation rate. Cubic spin-orbit coupling does not yield such a vanishing\nof the DP spin relaxation rate. Only in quantum dots whose siz e does not exceed the elastic mean free path lethe\nDP spin relaxation from cubic spin relaxation becomes dimin ished. In quantum wires , the electrons have a continuous\nspectrum of delocalized states. Still, transverse confinem ent can reduce the DP spin relaxation as we review in the\nnext section.\nIV. SPIN-DYNAMICS IN QUANTUM WIRES\nA. One-Dimensional Wires\nIn one dimensional wires, whose width Wis of the order of the Fermi wave length λF, impurities can only reverse the\nmomentum p→ −p. Therefore, the spin-orbit field can only change its sign, wh en a scattering from impurities occurs.\nBSO(p)→BSO(−p) =−BSO(p). Therefore, the precession axis and the amplitude of the spi n orbit field does not\nchange, reversing only the spin precession, so that the D’ya konov-Perel’-spin relaxation is absent in one dimensional\nwires.48In an external magnetic field, the precession around the magn etic field axis, due to the Zeeman-interaction is\ncompeting with the spin-orbit field, however. Then, as the el ectrons are scattered from impurities, both the precession\naxis and the amplitude of the total precession field is changi ng, since\n|B+BSO(−p)|=|B−BSO(p)|/negationslash=|B+BSO(p)|,\nresulting in spin dephasing and relaxation, as the sign of th e momentum changes randomly.\nB. Spin-Diffusion in Quantum Wires\nHow does the spin relaxation rate depend on the wire width Wwhen the quantum wire has more than one channel\noccupied,W >λ F? Clearly, for large wire widths, the spin relaxation rate sh ould converge to a finite value, while it\nvanishes for W→λF. It is both of practical importance for spintronic applicat ions and of fundamental interest to\nknow on which length scales this crossover occurs. Basicall y, there are three intrinsic length scales characterizing t he\nquantum wire relative to its width W. The Fermi wave length λF, the elastic mean free path leand the spin precession\nlengthLSO, Eq. ( 18). Suppression of spin relaxation for wire widths not exceed ing the elastic mean free path le, has\nbeen predicted and obtained numerically in Refs. [ 11,49–53]. Is the spin relaxation rate also suppressed in diffusive\nwires in which the elastic mean free path is smaller than the wire wi dth as in the wire shown schematically in Fig. 7?\nWe will answer this question by means of an analytical deriva tion in the following. The transversal confinement\nimposes that the spin current vanishes normal to the boundar y,jSin|Boundary= 0. For a wire grown along the [010]15\nFigure 7: Elastic scatterings from impurities and from the b oundary of the wire change the direction of the spin-orbit fie ld\naround which the electron spin is precessing.\ndirection, n= ˆexis the unit vector in the x-direction. For wire widths Wsmaller than the spin precession length LSO,\nthe solutions with the lowest energy have thus a vanishing tr ansverse spin current, and the spin diffusion equation\nEq. (20) becomes\n∂si\n∂t=−De∂yjSiy+τ/angb∇acketleft∇vF(BSO(k)×S)i/angb∇acket∇ight−/summationdisplay\nj1\nˆτsijsj. (41)\nwith\njSix|x=±W/2= (−τ/angb∇acketleftvx(BSO(k)×S)i/angb∇acket∇ight−De∂xSi)|x=±W/2= 0, (42)\nwhereWis the width of the wire. One sees that this equation has a pers istent solution, which does not decay in time\nand is homogeneous along the wire, ∂yS= 0. In this special case the spin diffusion equation simplifies t o12\n∂tS=−1\nτsα2\nα2\n1−α1α20\n−α1α2α2\n20\n0 0 α2\nS. (43)\nIndeed this has one persistent solution given by\nS=S0\nα2\nα1\n0\n, (44)\nThus, we can conclude that the boundary conditions impose an effective alignment of all spin-orbit fields, in a direction\nidentical to the one it would attain in a one-dimensional wir e, along the [010]-direction, setting kx= 0in Eq. ( 27),\nBSO(k) =−2ky\nα2\nα1\n0\n, (45)\nwhich therefore does not change its direction when the elect rons are scattered. This is remarkable, since this alignmen t\nalready occurs in wires with many channels, where the impuri ty scattering is two-dimensional, and the transverse\nmomentum kxactually can be finite. Rather, the alignment of the spin-orb it field, accompanied by a suppression of\nthe DP spin relaxation rate occurs due to the constraint on th e spin-dynamics imposed by the boundary conditions\nas soon as the wire width Wis smaller than the length scale which governs the spin dynam ics, namely, the spin\nprecession length LSO. It turns out that the spin diffusion equation Eq. ( 41) has also two long persisting spin helix\nsolutions in narrow wires13,54which oscillate periodically with the period LSO=π/m∗α. In contrast to the situation\nin 2D systems we reviewed in the previous Section, in quantum wires of width W < L SOthese solutions are long\npersisting even for α1/negationslash=α2. These two stationary solutions, are\nS=S0\nα1\nα\n−α2\nα\n0\nsin/parenleftbigg2π\nLSOy/parenrightbigg\n+S0\n0\n0\n1\ncos/parenleftbigg2π\nLSOy/parenrightbigg\n, (46)16\n0\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExtLSO\n2\nLSOy\n/Minus101\nSy/Slash1S0/Minus101\nSz/Slash1S0\n0/FractionBarExt/FractionBarExt/FractionBarExtΠ\n4/FractionBarExt/FractionBarExt/FractionBarExtΠ\n2/CurlyPhi\nFigure 8: Persistent spin helix solution of the spin diffusio n equation in a quantum wire whose width Wis smaller than the spin\nprecession length LSOfor varying ratio of linear Rashba α2=αsinϕand linear Dresselhaus coupling, α1=αcosϕ, Eq.( 46),\nfor fixed αandLSO=π/m∗α.\nand the linearly independent solution, obtained by interch angingcosandsinin Eq. ( 46). The spin precesses as the\nelectrons diffuse along the quantum wire with the period LSO, the spin precession length, forming a persistent spin\nhelix, whose x-component is proportional to the linear Dresselha us-coupling α1while its y-component is proportional\nto the Rashba-coupling α2as seen in Fig. 8. A similar reduction of the spin relaxation rate is not effect ive for cubic\nspin-orbit coupling for wire widths exceeding the elastic m ean free path le. One can derive the spin relaxation rate as\nfunction of the wire width for diffusive wires leL SOmodes which are localized at the boundaries and have a lower r elaxation rate than the bulk modes.12,13\nFor pure Rashba spin relaxation we find that there is a spin-he lix solution located at the edge whose relaxation rate\n1/τs=.31/τs0is smaller than the spin relaxation rate of bulk modes 1/τs= 7/16τs0.\nC. Weak Localization Corrections\nQuantum interference of electrons in low-dimensional, dis ordered conductors results in corrections to the electrica l\nconductivity ∆σ. This quantum correction, the weak localization effect, is k nown to be a very sensitive tool to study\ndephasing and symmetry breaking mechanisms in conductors.57–59The entanglement of spin and charge by spin-orbit\ninteraction reverses the effect of weak localization and the reby enhances the conductivity, the weak antilocalization\neffect. The quantum correction to the conductivity ∆σarises from the fact, that the quantum return probability\nto a given point x0after a time t,P(t), differs from the classical return probability, due to quant um interference.\nAs the electrons scatter from impurities, there is a finite pr obability that they diffuse on closed paths, which does\nincrease the lower the dimension of the conductor. Since an e lectron can move on such a closed orbit clockwise or\nanticlockwise as shown in light and dark blue in Fig. 9, with equal probability, the probability amplitudes of bot h\npaths add coherently, if their length is smaller than the dep hasing length Lϕ. In a magnetic field, indicated by the\nFigure 9: Electrons can diffuse on closed paths, orbit clockw ise or anticlockwise as indicated by the light and dark blue a rrows,\nrespectively. Middle figure: Closed electron paths enclose a magnetic flux from an external magnetic field, indicated as t he red\narrow, breaking time reversal symmetry. Right figure: The sc attering from a magnetic impurity spin, breaks the time reve rsal\nsymmetry between the clock- and anticlockwise electron pat hs.\nred arrow in the middle Fig. 9, the electrons acquire a magnetic flux phase. This phase depe nds on the direction in\nwhich the electron moves on the closed path. Thus, the quantu m interference is diminished in an external magnetic\nfield since the area of closed paths and thereby the flux phases are randomly distributed in a disordered wire, even\nthough the magnetic field can be constant. Similarly, the sca ttering from magnetic impurities breaks the time reversal\ninvariance between the two directions in which the closed pa th can be transversed. Therefore magnetic impurities\ndiminish the quantum corrections in proportion to the rate w ith which the electron spins scatter from them due to\nthe exchange interaction, 1/τMs, Eq. ( 38).\nThus, the quantum correction to the conductivity, ∆σis proportional to the integral over all times smaller than t he\ndephasing time τϕof the quantum mechanical return probability P(t) =λd\nFρ(x,t), wheredis dimension of diffusion,\nandρis the electron density. In the presence of spin-orbit scatt ering, the sign of the quantum correction changes to\nweak antilocalization as was as predicted by Hikami, Larkin , and Nagaoka60for conductors with impurities of heavy\nelements. As conduction electrons scatter from such impuri ties, the spin-orbit interaction randomizes their spin,\nFig.10. The resulting spin relaxation suppresses interference of time reversed paths in spin triplet configurations,\nwhile interference in singlet configuration remains unaffec ted as indicated in Fig. 10. Since singlet interference reduces\nthe electron’s return probability it enhances the conducti vity, the weak antilocalization effect. Weak magnetic fields\nsuppress also these singlet contributions, reducing the co nductivity and resulting in negative magnetoconductivity .\nIf the host lattice of the electrons provides spin-orbit int eraction, the spin relaxation of DP or EY type does have\nthe same effect of diminishing the quantum corrections in the triplet configuration. When the dephasing length\nLϕis smaller than the wire width W, the quantum corrections are determined by the interferenc e of 2-dimensional\nclosed diffusion paths, and as a result, the conductivity inc reases logarithmically with Lϕwhich increases itself as\nthe temperature is lowered. At low temperatures, the electr on-electron scattering is the dominating mechanism of\nspin dephasing, yielding Lϕ∼T−1/2. One can derive the magnetic field dependence of that quantum correction18\nFigure 10: As electrons diffuse, their spin precesses around the spin-orbit field, which changes its orientation, when th e electron\nis scattered. Electrons which enter closed paths with the sa me spin leave it therefore with a different spin if they choose the\npath in the opposite sense, as indicated by the light and dark blue arrows. However, electrons which enter the closed path with\nopposite spin, and move through the closed path in opposite s ense, attain the same quantum phase. This is a consequence of\ntime reversal invariance.\nnonperturbatively.42,60–64An approximate expression showing the logarithmic depende nce explicitly is given by\n∆σ=−1\n2πlnB+4\n3HMs+Hϕ\nHτ+1\n2πlnB+Hϕ+Hs+2\n3HMs\nHτ+1\nπlnB+Hϕ+cHs+2\n3HMs\nHτ, (50)\nin units of e2/h. All parameters are rescaled to dimensions of magnetic field s:Hϕ= 1/(4eDeτϕ) = 1/(4eL2\nϕ),\nHτ=/planckover2pi1/(4eDeτ), the spin relaxation field due to spin orbit relaxation, Hs=/planckover2pi1/(4eDeτs),61and the spin relaxation\nfield due to magnetic impurities HMs=/planckover2pi1/(4eDeτMs). Here1/τsis the DP relaxation rate in the 2D limit derived in\nthe previous section.61,65The prefactor cdepends on the particular spin-orbit interaction. For line ar Rashba-coupling,\nc= 7/16. Note that 7/16τsis the smallest spin relaxation rate of an inhomogeneous spi n density distribution13as\nderived in the section II B 4.1/τMsis the magnetic scattering rate from magnetic impurities, E q. (38). Indeed we\nsee that the first term does not depend on the DP spin relaxatio n rate. This term originates from the interference\nof time reversed paths, indicated in Fig. 10, which contributes to the quantum conductance in the single t state,\n|S= 0;m= 0/angb∇acket∇ight= (| ↑↓/angb∇acket∇ight− | ↓↑/angb∇acket∇ight )/√\n2, the minus sign in front of the second term is the origin of the change in sign\nin the weak localization correction. The other three terms a re suppressed by the spin relaxation rate, since they\noriginate from interference in triplet states, |S= 1;m= 0/angb∇acket∇ight= (| ↑↓/angb∇acket∇ight+| ↓↑/angb∇acket∇ight)/√\n2,|S= 1;m= 1/angb∇acket∇ight,|S= 1;m=−1/angb∇acket∇ight\nwhich do not conserve the spin symmetry. Thus, at strong spin -orbit induced spin relaxation the last three terms are\nsuppressed and the sign of the quantum correction switches t o weak antilocalization. In quasi-1-dimensional quantum\nwires which are coherent in transverse direction, W < L ϕthe weak localization correction is further enhanced, and\nincreases linearly with the dephasing length Lϕ. Thus, for WQSO≪1the weak localization correction is54\n∆σ=√HW/radicalBig\nHϕ+1\n4B∗(W)+2\n3HMs−√HW/radicalBig\nHϕ+1\n4B∗(W)+Hs(W)+2\n3HMs\n−2√HW/radicalBig\nHϕ+1\n4B∗(W)+1\n2Hs(W)+4\n3HMs, (51)\nin units ofe2/h. We defined HW=/planckover2pi1/(4eW2), and the effective external magnetic field,\nB∗(W) =/parenleftbigg\n1−1//parenleftbigg\n1+W2\n3l2\nB/parenrightbigg/parenrightbigg\nB. (52)19\nThe spin relaxation field Hs(W)is forW W, the spin relaxation rate is suppressed in analogy to the flux cancellation effect,\nwhich yields the weaker rate, 1/τs= (W/Cle)(DeW2/12L4\nSO), whereC= 10.8.66–68A dimensional crossover from\nweak antilocalization to weak localization and a reduction of spin relaxation has recently been observed experimental ly\nin quantum wires as we will review in the next Section.\nV. EXPERIMENTAL RESULTS ON SPIN RELAXATION RATE IN SEMICOND UCTOR QUANTUM\nWIRES\nA. Optical Measurements\nOptical time-resolved Faraday rotation (TRFR) spectrosco py69has been used to probe the spin dynamics in an\narray of n-doped InGaAs wires by Holleitner et al. in Ref. [ 70,71]. The wires were dry etched from a quantum well\ngrown in the [001]-direction with a distance of 1 µm between the wires. Spin aligned charge carriers were creat ed\nby absorption of circularly-polarized light. For normal in cidence, the spins point then perpendicular to the quantum\nwell plane, in the growth direction [001]. The time evolutio n of the spin polarization was then measured with a\nlinearly polarized pulse, see inset of Fig. 1c of Ref. [ 70]. The time dependence fits well with an exponential decay\n∼exp(−∆t/τs). As seen in Fig. 2a of Ref. [ 70], the thus measured lifetime τsat fixed temperature T= 5K of the\nspin polarization is enhanced when the wire width Wis reduced70: While for W >15µm it isτs= (12±1)ps,\nit increases for channels grown along the [100]- direction t o almostτs= 30 ps, and in the [110]- direction to about\nτs= 20ps. Thus, the experimental results show that the spin relaxa tion depends on the patterning direction of the\nwires: wires aligned along [100] and [010] show equivalent s pin relaxation times, which are generally longer than the\nspin relaxation times of wires patterned along [110] and [ 110]. The dimensional reduction could be seen already for\nwire widths as wide as 10µm, which is much wider than both the Fermi wave length and the e lastic mean free path\nlein the wires. This agrees well with the predicted reduction o f the DP scattering rate, Eq. ( 47) for wire widths\nsmaller than the spin precession length LSO. From the measured 2D spin diffusion length Ls(2D) = (0.9−1.1)µm,20\nand its relation to the spin precession length Eq. ( 18),LSO= 2πLs(2D), we expect the crossover to occur on a scale\nofLSO= (5.7−6.9)µm as observed in Fig. 2a of Ref. [ 70]. FromLSO=π/m∗αwe get with m∗= 0.064mea\nspin-orbit coupling α= (5−6)meVÅ. According to Ls=√Deτs, the spin relaxation length increases by a factor of/radicalbig\n30/12 = 1.6in the [100]-, and by/radicalbig\n20/12= 1.3in the [110]- direction.\nThe spin relaxation time has been found to attain a maximum, h owever, at about W= 1µm≈Ls(2D), decaying\nappreciably for smaller widths. While a saturation of τscould be expected according to Eq. ( 47) for diffusive wires,\ndue to cubic Dresselhaus-coupling, a decrease is unexpecte d. Schwab et al., Ref. [ 12], noted that with wire boundary\nconditions which do not conserve the spin of the conduction e lectrons one can obtain such a reduction. A mechanism\nfor such spin-flip processes at the edges of the wire has not ye t been identified, however. The magnetic field dependence\nof the spin relaxation rate yields further confirmation that the dominant spin relaxation mechanism in these wires is\nDPS: It follows the predicted behavior Eq. ( 40), as seen in Fig. 3a of Ref. [ 70], and the spin relaxation rate is enhanced\ntoτs(B= 1T) = 100 ps for all wire growth directions, at T= 5K and wire widths of W= 1.25µm.\nB. Transport Measurements\nA dimensional crossover from weak antilocalization to weak localization and a reduction of spin relaxation has\nrecently been observed experimentally in n-doped InGaAs qu antum wires,72,73in GaAs wires,74as well as in Al-\nGaN/GaN wires.75The crossover indeed occurred in all experiments on the leng th scale of the spin precession length\nLSO. We summarize in the following the main results of these expe riments.\nWirthmann et al., Ref. [ 72], measured the magnetoconductivity of inversion-doped In As quantum wells with a density\nofn= 9.7×1011/cm2, and a measured effective mass of m∗= 0.04me. In the wide wires the magnetoconductivity\nshowed a pronounced weak antilocalization peak, which agre ed well with the 2D theory,61,65with a spin-orbit-coupling\nparameter of α= 9.3meVÅ. They observed a diminishment of the antilocalization peak which occurred for wire widths\nW <0.6µm, atT= 2K, indicating a dimensional reduction of the DP spin relaxat ion rate.\nSchäpers et al. observed in Ga xIn1−xAs/InP quantum wires a complete crossover from weak antiloc alization to weak\nlocalization for wire widths below W= 500 nm. Such a crossover has also been observed in GaAs-quantum w ires by\nDinter et al., Ref. [ 74].\nVery recently, Kunihashi et al., Ref. [ 76] observed the crossover from weak antilocalization to weak localization in\ngate controlled InGaAs quantum wires. The asymmetric poten tial normal to the quantum well could be enhanced by\napplication of a negative gate voltage, yielding an increas e of the SIA-coupling parameter α, with decreasing carrier\ndensity, as was obtained by fitting the magnetoconductivity of the quantum wells to the theory of 2D weak localization\ncorrections of Iordanskii et al., Ref. [ 65]. Thereby, the spin relaxation length Ls=LSO/2πwas found to decrease from\n0.5µm to0.15µm, which according to LSO=π/m∗αcorresponds to an increase of αfrom(20±1)meVÅ at electron\nconcentrations of n= 1.4×1012/cm2toα= (60±1)meVÅ at electron concentrations of n= 0.3×1012/cm2. The\nmagnetoconductivity of a sample with 95 quantum wires in par allel showed a clear crossover from weak antilocaliza-\ntion to localization. Fitting the data to Eq. ( 51) a corresponding decrease of the spin relaxation rate was ob tained,\nwhich was observable already at large widths of the order of t he spin precession length LSOin agreement with the\ntheory Eq. ( 47). However, a saturation as obtained theoretically in diffus ive wires, due to cubic BIA-coupling was not\nobserved. This might be due to the limitation of Eq. ( 47), to diffusive wire widths, le< W, while in ballistic wires\na suppression also of the spin relaxation due to cubic BIA-co upling can be expected, since it vanishes identically in\n1-D wires, see section IVA. Also, an increase of the spin scattering rate in narrower wi res,W < L s(2D), was not\nobserved in contrast to the results of the optical experimen ts, Ref. [ 70], reviewed above.\nThe dimensional crossover has also been observed in the hete rostructures of the wide gap semiconductor GaN.75The\nmagnetoconductivity of 160 AlGaN/GaN- quantum wires were m easured. The effective mass is m∗= 0.22me, all\nwires were diffusive with le< W. For electron densities of n≈5×1012/cm2an increase from Ls(2D)≈550nm\ntoLs(W≈130nm)>1.8µm, and for densities n≈2×1012/cm2an increase from Ls(2D)≈500nm to\nLs(W≈120nm)>1. µm was observed. Using Ls(2D) = 1/2m∗α, one obtains for both densities n, the spin-\norbit coupling α≈5.8meVÅ. A saturation of the spin relaxation rate could not be ob served, suggesting that the\ncubic BIA-coupling is negligible in these structures.\nWe note, that an enhancement of the spin relaxation rate as in the optical experiments of narrow InGaAs quantum\nwires, Ref. [ 70], was not observed in these AlGaN/GaN-wires.\nVI. CRITICAL DISCUSSION AND FUTURE PERSPECTIVE\nThe fact that optical and transport measurements seem to find opposite behavior, enhancement and suppression\nof the spin relaxation rate, respectively, in narrow wires, calls for an extension of the theory to describe the crossove r21\nto ballistic quantum wires. This can be done, using the kinet ic equation approach to the spin-diffusion equation,12a\nsemiclassical approach,77,78or an extension of the diagrammatic approach.13In particular, the dimensional crossover\nof DPS due to cubic Dresselhaus coupling, which we found not t o be suppressed in diffusive wires, needs to be studied\nfor ballistic wires, le> W, as many of the experimentally studied quantum wires are in t his regime. Furthermore,\nusing the spin diffusion equation, one can study the dependen ce on the growth direction of quantum wires, and find\nmore information on the magnitude of the various spin-orbit coupling parameters, α1,α2,γD, by comparison with the\ndirectional dependence found in both the optical measureme nts70of the spin relaxation rate, as well as in recent gate\ncontrolled transport experiments.76\nIn narrow wires, corrections due to electron-electron inte raction can become more important and influence especially\nthe temperature dependence. Ref. [ 71] reports a strong temperature dependence of the spin relaxa tion rate in narrow\nquantum wires. As shown in Ref. [ 23], the spin relaxation rates obtained from the spin diffusion equation and the\nquantum corrections to the magnetoconductivity can be diffe rent, when corrections due to electron-electron interacti on\nbecome important. As the DPS becomes suppressed in quantum w ires other spin relaxation mechanisms like the EYS\nmay become dominant, since it is expected that the dimension al dependence of EYS is less strong. In more narrow\nwires, disorder can also result in Anderson localization. S imilar as in quantum dots,41,45this can yield enhanced spin\nrelaxation due to hyperfine coupling, Eq. ( 39). The spin relaxation in metal wires is believed to be domina ted by\nthe EYS mechanism, which is not expected to show such a strong wire width dependence, although this needs to be\nexplored in more detail. Even dilute concentrations of magn etic impurities of less than 1 ppm, do yield measurable\nspin relaxation rates in metals and allow the study of the Kon do effect with unprecedented accuracy.34,35\nVII. SUMMARY\nThe spin dynamics and spin relaxation of itinerant electron s in disordered quantum wires with spin-orbit coupling\nis governed by the spin diffusion equation Eq. ( 20). We have shown that it can be derived by using classical rand om\nwalk arguments, in agreement with more elaborate derivatio ns.12,13In semiconductor quantum wires all available\nexperiments show that the motional narrowing mechanism of s pin relaxation, the D’yakonov-Perel’-Spin relaxation\n(DPS) is the dominant mechanism in quantum wires whose width exceeds the spin precession length LSO. The solution\nof the spin diffusion equation reveals existence of persiste nt spin helix modes when the linear BIA- and the SIA-spin-\norbit coupling are of equal magnitude. In quantum wires whic h are more narrow than the spin precession length LSO\nthere is an effective alignment of the spin-orbit fields givin g rise to long living spin density modes for arbitrary ratio\nof the linear BIA- and the SIA-spin-orbit coupling. The resu lting reduction in the spin relaxation rate results in a\nchange in the sign of the quantum corrections to the conducti vity. Recent experimental results confirm the increase\nof the spin relaxation rate in wires whose width is smaller th anLSO, both the direct optical measurement of the spin\nrelaxation rate, as well as transport measurements. These s how a dimensional crossover from weak antilocalization\nto weak localization as the wire width is reduced. Open probl ems remain, in particular in narrower, ballistic wires,\nwere optical and transport measurements seem to find opposit e behavior of the spin relaxation rate: enhancement,\nsuppression, respectively. The experimentally observed r eduction of spin relaxation in quantum wires opens new\nperspectives for spintronic applications, since the spin- orbit coupling and therefore the spin precession length rem ains\nunaffected, allowing a better control of the itinerant elect ron spin. The observed directional dependence moreover\ncan yield more detailed information about the spin-orbit co upling, enhancing the spin control for future spintronic\ndevices further.22\nSymbols\nτ0elastic scattering time\nτeescattering time due to electron-electron interaction\nτepscattering time due to electron-phonon interaction\nτtotal scattering time 1/τ= 1/τ0+1/τee+1/τep.\nˆτsspin relaxation tensor\nDediffusion constant, De=v2\nFτ/dD, wheredDis the dimension of diffusion.\nleelastic mean free path\nLSOspin precession length in 2D. The spin will be oriented again in the initial direction after it moved ballistically\nthe lengthLSO.\nQSO= 2π/LSO\nLsspin relaxation length, Ls(W) =/radicalbig\nDeτs(W)withLs(W)|W→∞=Ls(2D) =LSO/2π\nLϕdephasing length\nα1linear (Bulk inversion Asymmetry (BIA) = Dresselhaus)-par ameter\nα2linear (Structural inversion Asymmetry (SIA) =Bychkov-Ra shba)-parameter\nγDcubic (Bulk inversion Asymmetry (BIA) = Dresselhaus)-para meter\nγggyromagnetic ratio\nAcknowledgments\nWe thank V. L. Fal’ko, F. E. Meijer, E. Mucciolo, I. Aleiner, C . Marcus, T. Ohtsuki, K. Slevin, J. Ohe, and A.\nWirthmann for helpful discussions. This research was suppo rted by DFG-SFB508 B9 and by WCU ( World Class\nUniversity ) program through the Korea Science and Engineer ing Foundation funded by the Ministry of Education,\nScience and Technology (Project No. R31-2008-000-10059-0 ).\n∗p.wenk@jacobs-university.de ;www.physnet.uni-hamburg.de/hp/pwenk/\n†s.kettemann@jacobs-university.de ;www.jacobs-university.de/ses/skettemann\n1I. Zutic, J. Fabian, and S. S. Das, Rev. Mod. Phys. 76, 323 (2004).\n2H. C. Torrey, Phys. Rev. 104, 563 (1956).\n3G. Dresselhaus, Phys. Rev. 100, 580 (1955).\n4E. Rashba, Soviet Physics-Solid State 2, 1109 (1960).\n5R. Lassnig, Phys. Rev. B 31, 8076 (1985).\n6R. Winkler, Spin-Orbit Coupling Effects in Two-Dimensional Electron an d Hole Systems , vol. 191 of Springer Tracts in\nModern Physics (Springer-Verlag, Berlin, 2003).\n7F. Malcher, G. Lommer, and U. Rössler, Superlattices and Mic rostructures 2, 267 (1986).\n8E. Bernardes, J. Schliemann, M. Lee, J. C. Egues, and D. Loss, Phys. Rev. Lett. 99, 076603 (2007).\n9J. Fabian, A. Matos-Abiague, C. Ertler, P. Stano, and I. Zuti c, Acta Physica Slovaca 57, 565 (2007).\n10S. Datta and B. Das, App. Phys. Lett. 56, 665 (1990).\n11A. G. Mal’shukov and K. A. Chao, Phys. Rev. B 61, R2413 (2000).\n12P. Schwab, M. Dzierzawa, C. Gorini, and R. Raimondi, Phys. Re v. B74, 155316 (2006).\n13P. Wenk and S. Kettemann, Phys. Rev. B 81, 18 (2010).\n14V. A. Froltsov, Phys. Rev. B 64, 045311 (2001).\n15N. Bloembergen, E. M. Purcell, and R. V. Pound, Phys. Rev. 73, 679 (1948).\n16M. I. D’yakonov and V. I. Perel’, Sov. Phys. Solid State 13(1972), [ Fiz. Tverd. Tela , 13:3581, 1971].\n17N. S. Averkiev and L. E. Golub, Phys. Rev. B 60, 15582 (1999).23\n18B. A. Bernevig, J. Orenstein, and S.-C. Zhang, Phys. Rev. Let t.97, 236601 (2006).\n19Y. Ohno, R. Terauchi, T. Adachi, F. Matsukura, and H. Ohno, Ph ys. Rev. Lett. 83, 4196 (1999).\n20C. P. Weber, J. Orenstein, B. A. Bernevig, S.-C. Zhang, J. Ste phens, and D. D. Awschalom, Phys. Rev. Lett. 98, 076604\n(2007).\n21M. M. Glazov and E. L. Ivchenko, Jetp Lett. 75, 403 (2002).\n22M. M. Glazov and E. L. Ivchenko, J. Exp. and Theoret. Phys. 99, 1279 (2004).\n23A. Punnoose and A. M. Finkel’stein, Phys. Rev. Lett. 96, 057202 (2006).\n24A. Dyson and B. K. Ridley, Phys. Rev. B 69, 125211 (2004).\n25R. J. Elliott, Phys. Rev. 96, 266 (1954).\n26Y. Yafet, in Solid State Physics , edited by F. Seitz and D. Turnbull (Academic, New York, 1963 ), vol. 14.\n27J. N. Chazalviel, Phys. Rev. B 11, 1555 (1975).\n28G. E. Pikus and A. N. Titkov, in Optical Orientation , edited by F. Meier and B. P. Zakharchenya (North-Holland, A mster-\ndam, 1984), vol. 8 of Modern Problems in Condensed Matter Sciences , chap. 3.\n29C. Grimaldi and P. Fulde, Phys. Rev. B 55, 15523 (1997).\n30P. Boguslawski, Solid State Commun. 33, 389 (1980).\n31G. L. Bir, A. G. Aronov, and G. E. Pikus, Sov. Phys.-JETP 42, 705 (1976), [ Zh. Eksp. Teor. Fiz. , 69:1382, 1975].\n32G. E. Pikus and G. L. Bir, Sov. Phys.-JETP 33, 108 (1971), [ Zh. Eksp. Teor. Fiz. , 60:195, 1971].\n33Müller-Hartmann, E., and J. Zittartz, Phys. Rev. Lett. 26, 428 (1971).\n34T. Micklitz, A. Altland, T. A. Costi, and A. Rosch, Phys. Rev. Lett.96, 226601 (2006).\n35G. Zaránd, L. Borda, J. von Delft, and N. Andrei, Phys. Rev. Le tt.93, 107204 (2004).\n36G. V. Astakhov, R. I. Dzhioev, K. V. Kavokin, V. L. Korenev, M. V. Lazarev, M. N. Tkachuk, Y. G. Kusrayev, T. Kiessling,\nW. Ossau, and L. W. Molenkamp, Phys. Rev. Lett. 101, 076602 (2008).\n37A. W. Overhauser, Phys. Rev. 89, 689 (1953).\n38E. L. Ivchenko, Sov. Phys. Solid State 15, 1048 (1973).\n39E. L. Ivchenko, Fiz. Tverd. Tela 15, 1566 (1973).\n40A. A. Burkov and L. Balents, Phys. Rev. B 69, 245312 (2004).\n41A. V. Khaetskii and Y. V. Nazarov, Phys. Rev. B 61, 12639 (2000).\n42I. L. Aleiner and V. I. Fal’ko, Phys. Rev. Lett. 87, 256801 (2001).\n43S. I. Erlingsson, Y. V. Nazarov, and V. I. Fal’ko, Phys. Rev. B 64, 195306 (2001).\n44A. Khaetskii, D. Loss, and L. Glazman, Phys. Rev. B 67, 195329 (2003).\n45A. V. Khaetskii, D. Loss, and L. Glazman, Phys. Rev. Lett. 88, 186802 (2002).\n46R. I. Dzhioev, K. V. Kavokin, V. L. Korenev, M. V. Lazarev, B. Y . Meltser, M. N. Stepanova, B. P. Zakharchenya,\nD. Gammon, and D. S. Katzer, Phys. Rev. B 66, 245204 (2002).\n47P. I. Tamborenea, D. Weinmann, and R. A. Jalabert, Phys. Rev. B76, 085209 (2007).\n48J. S. Meyer, V. I. Fal’ko, and B. L. Altshuler (Kluwer Academi c Publishers, Dordrecht, 2002), vol. 72 of Nato Science Series\nII, p. 117.\n49A. Bournel, P. Dollfus, P. Bruno, and P. Hesto, The European P hysical Journal Applied Physics 4, 1 (1998).\n50A. A. Kiselev and K. W. Kim, Phys. Rev. B 61, 13115 (2000).\n51T. P. Pareek and P. Bruno, Phys. Rev. B 65, 241305 (2002).\n52T. Kaneko, M. Koshino, and T. Ando, Phys. Rev. B 78, 245303 (2008).\n53R. L. Dragomirova and B. K. Nikolic, Phys. Rev. B 75, 085328 (2007).\n54S. Kettemann, Phys. Rev. Lett. 98, 176808 (2007).\n55V. L. Fal’ko (2003), private communication.\n56T. Schäpers, V. A. Guzenko, M. G. Pala, U. Zülicke, M. Governa le, J. Knobbe, and H. Hardtdegen, Phys. Rev. B 74, 081301\n(2006).\n57B. L. Altshuler, A. G. Aronov, D. E. Khmelnitskii, and A. I. La rkin (Mir, Moscow, 1982), Quantum Theory of Solids.\n58G. Bergmann, Physics Reports 107, 1 (1984).\n59S. Chakravarty and A. Schmid, Physics Reports 140, 193 (1986).\n60S. Hikami, A. I. Larkin, and Y. Nagaoka, Progress of Theoreti cal Physics 63, 707 (1980).\n61W. Knap, C. Skierbiszewski, A. Zduniak, E. Litwin-Staszews ka, D. Bertho, F. Kobbi, J. L. Robert, G. E. Pikus, F. G. Pikus,\nS. V. Iordanskii, et al., Phys. Rev. B 53, 3912 (1996).\n62J. B. Miller, D. M. Zumbühl, C. M. Marcus, Y. B. Lyanda-Geller , D. Goldhaber-Gordon, K. Campman, and A. C. Gossard,\nPhys. Rev. Lett. 90, 076807 (2003).\n63Y. Lyanda-Geller, Phys. Rev. Lett. 80, 4273 (1998).\n64L. E. Golub, Phys. Rev. B 71, 235310 (2005).\n65S. V. Iordanskii, Y. Lyanda-Geller, and G. E. Pikus, JETP Let t.60, 206 (1994).\n66C. W. J. Beenakker and H. van Houten, Phys. Rev. B 37, 6544 (1988).\n67V. K. Dugaev and D. E. Khmel’nitskii, Soviet Physics - JETP 59, 1038 (1984).\n68S. Kettemann and R. Mazzarello, Phys. Rev. B 65, 085318 (2002).\n69D. Stich, J. H. Jiang, T. Korn, R. Schulz, D. Schuh, W. Wegsche ider, M. W. Wu, and C. Schüller, Phys. Rev. B 76, 073309\n(2007).\n70A. W. Holleitner, V. Sih, R. C. Myers, A. C. Gossard, and D. D. A wschalom, Phys. Rev. Lett. 97, 036805 (2006).\n71A. W. Holleitner, V. Sih, R. C. Myers, A. C. Gossard, and D. D. A wschalom, New Jounral of Physics 9, 342 (2007).\n72A. Wirthmann, Y. Gui, C. Zehnder, D. Heitmann, C.-M. Hu, and S . Kettemann, Physica E: Low-dimensional Systems24\nand Nanostructures 34, 493 (2006), proceedings of the 16th International Confere nce on Electronic Properties of Two-\nDimensional Systems (EP2DS-16).\n73F. E. Meijer (2005), private communication.\n74R. Dinter, S. Löhr, S. Schulz, C. Heyn, and W. Hansen (2005), u npublished.\n75P. Lehnen, T. Schäpers, N. Kaluza, N. Thillosen, and H. Hardt degen, Phys. Rev. B 76, 205307 (2007).\n76Y. Kunihashi, M. Kohda, and J. Nitta, Phys. Rev. Lett. 102, 226601 (2009).\n77O. Zaitsev, D. Frustaglia, and R. K., Phys. Rev. Lett. 94, 026809 (2005).\n78O. Zaitsev, D. Frustaglia, and R. K., Phys. Rev. B 72, 155325 (2005)." }, { "title": "2011.02369v2.Dynamical_spin_correlations_of_the_kagome_antiferromagnet.pdf", "content": "Dynamical spin correlations of the kagome antiferromagnet\nP. Prelov\u0014 sek,1, 2M. Gomil\u0014 sek,1T. Arh,1, 2and A. Zorko1, 2,\u0003\n1Jo\u0014 zef Stefan Institute, Jamova c. 39, SI-1000 Ljubljana, Slovenia\n2Faculty of Mathematics and Physics, University of Ljubljana, Jadranska u. 19, SI-1000 Ljubljana, Slovenia\nTemperature-dependent dynamical spin correlations, which can be readily accessed via a variety of\nexperimental techniques, hold the potential of o\u000bering a unique \fngerprint of quantum spin liquids\nand other intriguing dynamical states. In this work we present an in-depth study of the temperature-\ndependent dynamical spin structure factor S(q;!) of the antiferromagnetic (AFM) Heisenberg spin-\n1/2 model on the kagome lattice with additional Dzyaloshinskii{Moriya (DM) interactions. Using\nthe \fnite-temperature Lanczos method on lattices with up to N= 30 sites we \fnd that even without\nDM interactions, chiral low-energy spin \ructuations of the 120 °AFM order parameter dominate the\ndynamical response. This leads to a nontrivial frequency dependence of S(q;!) and the appearance\nof a pronounced low-frequency mode at the M point of the extended Brillouin zone. Adding an\nout-of-plane DM interactions Dzgives rise to an anisotropic dynamical response, a softening of\nin-plane spin \ructuations, and, ultimately, the onset of a coplanar AFM ground-state order at\nDz>0:1J. Our results are in very good agreement with existing inelastic neutron scattering and\ntemperature-dependent NMR spin-lattice relaxation rate (1 =T1) data on the paradigmatic kagome\nAFM herbertsmithite, where the e\u000bect of its small Dzon the dynamical spin correlations is shown\nto be rather small, as well as with 1 =T1data on the novel kagome AFM YCu 3(OH) 6Cl3, where its\nsubstantial Dz\u00190:25Jinteraction is found to strongly a\u000bect the spin dynamics.\nI. INTRODUCTION\nThe antiferromagnetic (AFM) Heisenberg spin-1/2\nmodel on the kagome lattice (KLHM) is one of the\nmost intensively studied quantum spin models, owing\nto its unique ground state (GS) and low- Tproperties\n[1{4]. Various theoretical and numerical investigations\nhave established KLHM as the most promising candidate\namongst isotropic spin models to feature a quantum spin\nliquid (SL) GS, where the absence of low- Tlong-range\norder is accompanied by strong quantum entanglement\nbetween constituent spins. However, the nature of the\nSL GS, including the presence of either a \fnite [5{15] or\na vanishing [16{21] energy gap \u0001 tto spin-triplet excita-\ntions, remains controversial. Properties of the KLHM at\n\fnite temperatures may provide important insights into\nthis long-standing issue.\nThermodynamic quantities such as the uniform suscep-\ntibility\u001f0(T), magnetic speci\fc heat c(T), and the re-\nlated entropy density s(T) of the KLHM have previously\nbeen studied by high- Tseries expansion [22, 23], via nu-\nmerical linked cluster methods [24, 25], and more recently\nwith the \fnite-temperature Lanczos method (FTLM)\n[13, 15, 26] on \fnite spin systems with up to N= 42 sites.\nApart from an evidence of \fnite spin triplet gap \u0001 t>0,\nFTLM results indicate that there is a substantial rem-\nnant entropy s(T)>0 at very low T, which is a signature\nof a large density of low-energy singlet excitations with a\n(nearly) vanishing spin singlet energy gap \u0001 s\u001c\u0001t. The\nstatic (equal-time) spin correlation function S\u000b\u000b(q) has\nalso been studied both at T= 0 [17] and at \fnite tem-\nperatures [27, 28]. However, dynamical spin properties\n\u0003andrej.zorko@ijs.siof the KLHM, in particular the dynamical spin struc-\nture factor (DSF) S\u000b\u000b(q;!), are theoretically poorly un-\nderstood even though the temperature-dependent DSF\nis potentially a unique \fngerprint of SL states, and is\nexperimentally directly accessible via inelastic neutron-\nscattering (INS) and nuclear magnetic resonance (NMR)\nrelaxation [29]. Because of its fundamental importance\nvarious analytical concepts and methods [30{32], as well\nas numerical approaches [28, 33], have been employed to\nstudy it, though they have mostly led to inconclusive re-\nsults.\nOne reason for the theoretical di\u000eculties lies in the\nlarge density of low-energy spin-singlet states of the\nKLHM [15], which implies that a meaningful evalua-\ntion of the DSF would require a more challenging \fnite-\ntemperature instead of GS treatment. Another reason is\nthat the DSF of a SL, like the one in KLHM, is usually\n(implicitly) assumed to be rather featureless due to the\nfractionalization of spin excitations. We show that the\nKLHM DSF instead has some quite pronounced spectral\nfeatures.\nOn the experimental front, investigations of the KLHM\nhave been boosted in the last couple of decades by\nthe discovery of several promising kagome-lattice (KL)\nmaterials exhibiting SL properties at low tempera-\ntures. The most prominent example is herbertsmithite,\nZnCu 3(OH) 6Cl2[34{36], where the availability of single\ncrystals allows for full access to the DSF S\u000b\u000b(q;!) [37{\n39]. While several other KL materials have been discov-\nered in recent years [40{45], we will mostly focus on the\nrecently synthesized [46] and investigated [26, 45, 47, 48]\ncompound YCu 3(OH) 6Cl3, which has the distinct ad-\nvantage of having a structurally-perfect kagome lattice\nwithout any substitutional disorder, in contrast to most\nother KL materials including herbertsmithite [35, 36, 49].\nBesides potential imperfections the relation of KL ma-arXiv:2011.02369v2 [cond-mat.str-el] 20 Jan 20212\nterials to the ideal KLHM is often further complicated\nby additional Dzyaloshinskii{Moriya (DM) interactions,\nwhich are usually allowed in these systems as most lack\nlocal inversion symmetry on superexchange bonds Jbe-\ntween nearest-neighbor magnetic ions. While weak DM\ninteractions are expected to lead to mostly quantitative\ncorrections of observables at low T[23, 25], as in the\ncase of herbertsmithite [50, 51], strong DM interactions\ncan lead to a quantum phase transition from a SL to\na long-range ordered (LRO) GS [52{54], as in the case\nof YCu 3(OH) 6Cl3where an out-of-plane Dz\u00190:25Jin-\nduces chiral 120 °AFM LRO [26, 48]. The addition of\nDM interactions to the KLHM is therefore crucial for ex-\nplaining the observed properties of many KL materials,\nespecially low- Tordered ones like YCu 3(OH) 6Cl3.\nIn this paper we present a comprehensive numerical\nstudy of the DSF S\u000b\u000b(q;!) of the KLHM with additional\nout-of-plane DM interactions D=Dzat \fnite tempera-\ntures. To this end we employ the FTLM on systems with\nup toN= 30 sites under periodic boundary conditions.\nThis method is particularly suitable for frustrated spin\nsystems (and in general strongly-correlated systems) that\ndo not possess long-range correlations down to T\u001cJ,\nwhich allows us to obtain static and dynamical properties\nof macroscopic validity down to temperatures many times\nlower [15, 26] than in systems with GS LRO [15, 55]. In\ncontrast to previous investigations of the KLHM DSF [28]\nwe \fnd that it is in fact not featureless. Even at D= 0 we\n\fnd particularly pronounced low-energy chiral 120 °AFM\n\ructuations corresponding to the wavevector q= 0 in the\nreduced Brillouin zone (BZ), or, equivalently, the M point\nof the extended BZ. Furthermore, we \fnd that the low- T,\nlow-energy DSF of the KLHM seems to be governed by a\n\fnite spin triplet gap \u0001 t>0. Adding \fnite DM interac-\ntionsD> 0 results in an anisotropic DSF and a softening\nof the in-plane spin triplet gap \u0001x\ntthat ultimately leads\nto GS LRO for D > Dc\u00190:1J. The calculated DSF is\nalso used to evaluate temperature-dependent local spin\n\ructuation (LSF) spectra S\u000b\u000b\nL(!), which are directly re-\nlated to experimental NMR spin-lattice relaxation rates\n1=T1. Finally, the obtained numerical DSF and LSF re-\nsults are compared with experimental INS [37] and NMR\nresults [38] on herbertsmithite and on the impurity-free\nYCu 3(OH) 6Cl3[56].\nII. MODEL, NUMERICAL METHOD AND\nCONSIDERED QUANTITIES\nWe consider the KLHM with AFM isotropic Heisen-\nberg nearest-neighbor exchange interactions Jbetween\nS= 1=2 spins on a KL with additional out-of-plane DM\ninteractions D,\nH=X\nhijih\nJSi\u0001Sj+D(Si\u0002Sj)zi\n; (1)\nwherehijiis a sum over nearest-neighbor spin pairs and\nthe spins in the DM term appear in the clockwise direc-\nFigure 1. Finite size kagome lattices with N= 24, 27 and 30\nsites used in our FTLM calculations. The primitive vectors\nof the underlying hexagonal Bravais lattice are denoted by a1\nanda2, while the three basis vectors of the kagome lattice are\ndenoted by r0,r1, and r2.\ntion around each lattice triangle (see Fig. 1). Except in\nSec. V where we compare our numerical results with ex-\nperiment, we use ~=kB= 1 units as well as J= 1. All\nenergies, frequencies and temperatures are thus implicitly\ngiven relative to J. In KL materials the DM interaction\nhas the general form Dij\u0001(Si\u0002Sj) where Dijis a vec-\ntor with an out-of-plane component Dz\nijand an in-plane\ncomponent Dp\nij. In this paper we consider only the e\u000bect\nof a non-zero Dz\nij=Dfor three reasons. Firstly, an in-\nplaneDp\nijis symmetry-allowed only when the kagome\nplane is not also a crystallographic mirror plane [53],\nand is thus present less often. Secondly, the e\u000bect of\nDp\nij6= 0 appears to be weaker and qualitatively less im-\nportant than that of Dz\nij6= 0 in the KLHM, as con\frmed\nboth theoretically and experimentally [23, 25, 26, 52].\nAnd thirdly, as a practical bene\ft, when Dp\nij= 0 the\nhamiltonianHremains uniaxially symmetric about the\nzaxis, conserving the zcomponent of total magnetiza-\ntionSz\ntot=P\niSz\ni, which signi\fcantly reduces the di-\nmensionality of invariant Hilbert subspaces, and hence\nthe memory requirements, of the FTLM.\nThe standard de\fnition of the DSF is\nS\u000b\f(q;!) =1\n2\u0019Z1\n\u00001dt ei!thS\u000b\n\u0000q(t)S\f\nq(0)i;(2)\nwhereh:::idenotes the canonical thermal average, \u000b\nand\fare components of q-space spin operators Sq=\n(1=p\nN)P\nieiq\u0001RiSide\fned via the positions Riof spins\nin the KL, and Nis the total number of KL sites. As\nthe KL is formed of three basis vectors rk(k= 0;1;2) on\nan underlying hexagonal Bravais lattice of down-pointing\ntriangle centers eRn(see Fig. 1) one has Ri=eRn+rk\nwherei\u0011(n;k). Due to the three KL basis vectors\nthe DSF is only q-periodic over an extended BZ that\nis 4-times larger than the reduced BZ of the underlying\nhexagonal Bravais lattice (see Fig. 5).\nA more insightful de\fnition of the DSF for the KL,3\nwhich explicitly takes into account its threefold rota-\ntional symmetry, and which is bene\fcial both numerically\nas well as for theoretical understanding, instead involves\nchiral spin operators in q-space,\neScq=1p\nNX\nneiq\u0001eRn\u0002\nS(n;0)+\u0010cS(n;1)+\u0010\u0000cS(n;2)\u0003\n;\n(3)\nwherenin (n;k) runs over all down-pointing triangles of\nthe KL,kruns over the three spins inside these triangles,\n\u0010=e2\u0019i=3, andc=\u00001;0;1 denotes the vector spin chi-\nrality of the KL triangles. Note that the standard 120 °\nAFM LRO on the KL involves only the chiral spin oper-\nators withc=\u00061, while ferromagnetic LRO involves the\nc= 0 chiral spin operators. Using Eq. (3) we de\fne the\n(diagonal) chiral DSF on the KL as\neS\u000b\f\nc(q;!) =1\n2\u0019Z1\n\u00001dt ei!theS\u000by\ncq(t)eS\f\ncq(0)i;(4)\nwhich is q-periodic over the reduced BZ, not just over\nthe larger extended BZ of the standard DSF of Eq. (2)\n(see Fig. 5). Note that at a generic qone could also\nexpect non-vanishing o\u000b-diagonal terms heS\u000by\ncq(t)eS\f\nc0q(0)i\nwithc6=c0. Nevertheless, these terms are expected to\nbe less important than the diagonal ones, and are much\nmore di\u000ecult to handle within the FTLM, so we neglect\nthem. We can then express the standard DSF of Eq. (2)\nfrom the chiral DSF of Eq. (4) as\nS\u000b\u000b(q;!) =1X\nc=\u00001j\u0018c(q)j2eS\u000b\u000b\nc(q;!);\n\u0018c(q) =1\n32X\nk=0eiq\u0001rk\u0010\u0000ck;(5)\nwhereeS\u000b\f\nc(q;!) = 0 for \u000b6=\fandSxx(q;!) =\nSyy(q;!) sinceSz\ntotis a conserved quantity.\nWe evaluate the chiral DSF at T >0 using the FTLM,\nintroduced in Refs. [55, 57] and used in numerous stud-\nies of static and dynamical properties of various corre-\nlated systems [58]. In the case of the KLHM, the FTLM\nhas previously been employed only for the calculation of\nthermodynamic quantities, such as the uniform suscepti-\nbility\u001f0(T), entropy density s(T) and speci\fc heat c(T)\n[13, 15, 26], that involve only the conserved quantities\nof energy and total magnetization Sz\ntot. In contrast, the\nevaluation of the chiral DSF (here given in the Lehmann\nrepresentation) is more involved,\neS\u000b\u000b\nc(q;!) =1\nZX\nne\u0000\u000fn=Th njeS\u000by\ncq\u000e(!+\u000fn\u0000H)eS\u000b\ncqj ni;\n(6)\nwhereZ=P\nne\u0000\u000fn=Tis the canonical partition function,\nj niare eigenfunctions of Hand\u000fnare their eigenener-\ngies. As the chiral DSF already takes into account both\ntranslation symmetry and the conservation of Sz\ntot, the\nneeded Hilbert subspaces remain the same as for staticquantities; e.g., the largest subspace for N= 30 sites\ncontainsNst\u0018107states. In FTLM we replaceP\nnover\nall eigenfunctions with a trace over R> 1 random initial\nwavefunctionsjriand the expectation value with a dou-\nble sum over the emerging Lanczos (eigen)functions j\u001er\nii,\nje\u001er\njiin di\u000berent qwavevector sectors [55, 57, 58], with\ni;j\u0014NLwhereNLis the number of performed Lanczos\nsteps. This requires additional storage of 2 NLwavefunc-\ntions meaning that the total memory requirements for\nthe dynamical FTLM are O(NLNst). To achieve satis-\nfactory!resolution in the DSF NL>100 is typically\nrequired.\nIn the following we evaluate the chiral DSF on several\n\fnite-sized lattices with N= 24, 27 and 30 sites (Fig. 1).\nWhile theN= 24 and 30 lattices break the rotational\nsymmetry of the in\fnite KL the N= 27 lattice preserves\nit, but is less convenient because of its Stot= 1=2 GS,\nwhereas the in\fnite KLHM should have a Stot= 0 GS\n[1{4]. While for N= 24 and 27 we can a\u000bord NL\u0018200\nandR > 10, most of the present results are for N= 30\nsites where we used NL= 120 and R= 3 within each\nsymmetry sector. We note that the main criterion for\n(even macroscopic) validity of FTLM results (in the given\nmodel and system size) is that the modi\fed thermody-\nnamic sumeZ(T) =RTr[exp(\u0000(H\u0000E0)=T)]>eZ(Tfs)\u001d\n1 [55, 57], where E0is the ground-state energy and trace\nTr involves the sum over all wave vector and Sz\ntotsectors.\nDue to very large density of low-lying states in SL sys-\ntems (and directly related large entropy even at low T),\neven modest R= 3 is enough to reach valid results down\nto temperatures T >T fs\u00180:1J+D[15, 26] below which\nthey are limited by \fnite-size e\u000bects, i.e., by the onset of\nlonger-range correlations for D> 0.\nFinally, while S\u000b\u000b(q;!) contains all of the dynamical\ninformation, it is also useful to extract the equal-time\nspin correlation function S\u000b\u000b(q) and the d.c. spin sus-\nceptibility\u001f\u000b\u000b\n0(q), de\fned from the DSF as\nS\u000b\u000b(q) =Z1\n\u00001d! S\u000b\u000b(q;!) =hS\u000b\n\u0000qS\u000b\nqi;\n\u001f\u000b\u000b\n0(q) =PZ1\n\u00001d!1\u0000e\u0000!=T\n!S\u000b\u000b(q;!);(7)\nwherePdenotes the Cauchy principal value. Note that\n!\u000b(q) =p\nS\u000b\u000b(q)=\u001f\u000b\u000b\n0(q) can be interpreted as the\ncharacteristic spin-\ructuation frequency at a given qand\nfor a given direction \u000b.\nIII. HEISENBERG MODEL ON KAGOME\nLATTICE\nIn this section we consider the D= 0 KLHM. Since\nthis model is isotropic in spin space it has an isotropic\nchiral DSFeS\u000b\u000b\nc(q;!) =eSc(q;!) and hence also isotropic\nderived quantities in Eqs. (7). In the following we present\nnumerical results for the standard choice \u000b=z, which4\n-10 1 2 3 4 5 6 0.00.10.20.3 \nSc(q, ω)ω\n Γ, c = ±1 (N = 24) \nΓ, c = ±1 (N = 27) \nΓ, c = ±1 (N = 30) \nM, c = 0 (N = 24) \nM, c = 0 (N = 30)\nFigure 2. Chiral DSF's eS\u00061(q= 0;!) andeS0(q= M;!) at\nT= 0:2 calculated using the FTLM for di\u000berent lattice sizes\nN= 24{30 (Fig. 1). Note that the N= 27 lattice does not\ncontain the M point of the reduced BZ. The vertical dotted\nline at!= 1:5 corresponds to triplet excitations within an\nisolated Heisenberg spin triangle.\nis numerically less costly to evaluate since the relevant\noperators are diagonal in the Sz\ntotbasis.\nA. Dynamical spin structure factor\nIn Fig. 2 we present a comparison of chiral DSF's cal-\nculated at T= 0:2 using the FTLM on lattices with\nN= 24, 27 and 30 sites (see Fig. 1). We choose two\nrather extreme cases of the \u0000 and M points of the re-\nduced BZ [see inset in Fig. 3(b)]. The largest dynamical\nresponse is at the \u0000 point ( q= 0) with chirality c=\u00061,\nwhich represents uniform \ructuations of the AFM order\nparameter for 120 °ordered spins on each KL triangle.\nWe see that these results are quite independent of lattice\nsizeN, which con\frms that the spin correlation length\nis quite short even at this low temperature due to strong\ngeometric frustration. The spectra are not featureless, as\nthey exhibit two distinct frequency maxima, which seem\nquite robust. These were already tentatively observed via\nthe numerical linked cluster method [29] by assuming an\nad hoc Lorentzian line shape. Our FTLM calculations,\non the other hand, do not require any a priori assump-\ntions on the line shape. The higher-energy maximum can\nbe traced back to transitions within individual spin tri-\nangles, for which the energy gap between the S= 1=2\nGS and excited S= 3=2 spin states is != 1:5 (dashed\nline in Fig. 2).\nIn Fig. 3 we show the full chiral DSF eSc(q;!) atT=\n0:2 for all inequivalent q's in the reduced BZ for both\nchirality branches c=\u00061 andc= 0, calculated on the\nlargestN= 30 site lattice. Since the c= +1 andc=\u00001\nchiral DSF's are in general not equal at generic q6= 0 we\nplot in Fig. 3(a) the averaged chiral DSF\nS1(q;!) =1\n2h\neS1(q;!) +eS\u00001(q;!)i\n; (8)\n0.00.10.20.3\n-1 0 1 2 3 4 5 60.000.020.040.060.08\n(a)S1(q, ω)\n Γ\n q1\n q2\n q3\n M\n q5S0(q, ω)\nω(b)\nΓM\nq-1q1\nq2q-2\nq3q-3\nq-5\nq5Figure 3. (a) Average chiral DSF S1(q;!) [Eq. (8)], and (b)\nchiral DSF eS0(q;!) atT= 0:2 for all inequivalent numerical\nq's within the reduced BZ calculated on the N= 30 lattice.\nNote the very di\u000berent vertical scales of both panels. Inset\nin (b) shows the numerical q-cells in the reduced BZ of the\nN= 30 lattice (Fig. 1).\nwhile atq= 0 both chiralities c=\u00061 match and we have\nS1(q= 0;!) =eS\u00061(q= 0;!). We see that chiral c=\u00061\n\ructuations indeed dominate the response (Fig. 3), with\nthe largest intensity found at the q= 0 (\u0000) point and a\nslightly reduced intensity found at the smallest nonzero\nq=q1. Chiral DSF spectra at these low qshow the char-\nacteristic double-maximum frequency dependence with\nmaxima near !\u00190:3 and!\u00191:5 [Fig. 3(a)]. This struc-\nture is reproduced even in the considerably weaker c= 0\nresponse at q=q1[Fig. 3(b)]. At larger q, nearer the\nBZ boundary, all spectra are broad ( \u000e!&3), weak and\nfeatureless. The observed qandcdependence thus in-\ndicates that longer-ranged chiral 120 °AFM correlations\ndominate the dynamical response of the KLHM at low\nfrequencies, with a correlation length \u0018 > 1 extending\nfurther than a single spin triangle.\nIn Fig. 4 we present the temperature evolution of the\ndominantq= 0,c=\u00061 chiral DSF. It is evident that\nthe double-maximum frequency structure is not just a\nlow-Tfeature as it persists to temperatures as large as\nT\u00181. Partly, the low-energy peak at !\u00180:3 is simply\na consequence of the detailed balance relation for DSF's,\nS(\u0000!) = exp(\u0000!=T)S(!), which implies d S=d!j!=0=\nS(0)=(2T)>0 and thus always leads to a maximum at\n! > 0. On the other hand, at the lowest T= 0:1\u0018Tfs\nwe \fnd a further reduced != 0 response, which could\nindicate a \fnite spin triplet gap \u0001 t>0, at least on our\n\fnite-sized N= 30 lattice.\nFinally, we note that the q= 0,c= 0 chiral DSF had to5\nbe explicitly excluded from our FTLM calculations since\nit is singular in \fnite systems, eS0(q= 0;!)/\u000e(!), due\nto the conservation of Sz\ntot. Nevertheless, in the macro-\nscopic limit N!1 atT > 0 this singular DSF should\nevolve into the q\u00190 spin di\u000busion peak with a spectral\nwidth that is expected to scale as \u000e!/Ddi\u000bq2where\nDdi\u000b(T) is the temperature-dependent spin di\u000busion con-\nstant [59]. We discuss the experimental relevance of this\ncontribution in more detail in Sec. V B.\nB. Equal-time correlations and static response\nFor comparison with experimental INS data as well as\nwith previous theoretical calculations, it is informative\nto also look at the standard INS DSF S(q;!) [Eq. (2)] in\nthe extended BZ, which is calculated from the chiral DSF\neSc(q;!) via Eq. (5). Firstly, we consider the q-dependent\nequal-time spin correlation function S(q) [Eqs. (7)] over a\nbroad range of temperatures T= 0:1{2:0 on theN= 30\nlattice (Fig. 5). Consistent with several previous numer-\nical studies of this quantity [17, 27, 28], S(q) has a pro-\nnounced but spread-out region of high intensity around\nthe whole extended BZ boundary that remains visible\neven at very high T\u00182. This can be understood by\nconsidering the q-dependence of the chiral weighing fac-\ntorj\u0018c(q)j2in Eq. (5) that suppresses the contribution of\nthe dominant chiral c=\u00061 \ructuations to the standard\nDSFS(q;!) near the \u0000 point of the extended BZ, but\nnot at the extended BZ boundary. Weak global maxima\nofS(q) appear for T >0:1 at corner K points of the ex-\ntended BZ (note that our N= 30 lattice does not contain\nthis point), qualitatively consistent with previous stud-\nies, but appreciable intensity can also be found at the M\npoints (corresponding to periodic images of the \u0000 point\nof the reduced BZ).\nA complementary quantity, which is more sensitive to\nlow-energy \ructuations as is obvious from Eqs. (7), is the\nq-dependent d.c. susceptibility \u001f0(q), which we present\nin the extended BZ over a broad range of temperatures\n-2-101234560.00.10.20.3 T = 0.1 \nT = 0.2 \nT = 0.4 \nT = 0.8 \nT = 2 \nS±1(q = 0, ω)ω\nFigure 4. The chiral DSF eS\u00061(q= 0;!) at di\u000berent T= 0:1{\n2:0 on theN= 30 lattice.\nFigure 5. The equal-time spin correlation function S(q) in\nthe extended BZ (large hexagon) at di\u000berent T= 0:1{2:0 on\ntheN= 30 lattice. The reduced BZ is shown by the smaller\nhexagon. The left halves of the panels show raw FTLM results\nonq-cells shown in the inset in Fig. 3(b), while the right\nhalves are rotationally symmetrized (to recover the threefold\nsymmetry of the KL) and smoothed via interpolation.\nFigure 6. The d.c. spin susceptibility \u001f0(q) in the extended\nBZ (large hexagon) at di\u000berent T= 0:1{2:0 on theN= 30\nlattice. The reduced BZ is shown by the smaller hexagon.\nThe left halves of the panels show raw FTLM results on q-\ncells shown in the inset in Fig. 3(b), while the right halves are\nrotationally symmetrized as in Fig. 5.\nin Fig. 6. A striking di\u000berence to S(q) (Fig. 5) is a very\npronounced maximum of \u001f0(q) at the M point of the\nextended BZ, which is directly related to the dominant6\n-2-101234560.000.020.040.06 T = 0.1 \nT = 0.2 \nT = 0.4 \nT = 0.8 \nT = 2 \nSL(ω)ω\nFigure 7. The LSF SL(!) at di\u000berent T= 0:1{2:0 on the\nN= 30 lattice.\nlow-energy q= 0,c=\u00061 chiral \ructuations seen in\nthe chiral DSF (Fig. 3). This maximum is much more\nsensitive to temperature than the maximum in S(q) and\ndisappears for T > 1, consistent with the broadening of\nthe chiral response visible in Fig. 4. It should be stressed\nthat the same maximum is directly related to the one\nobserved by low-energy INS in herbertsmithite [37], as\nwill be discussed in more detail in Sec. V A.\nC. Local spin \ructuations\nLSF can be expressed from the chiral DSF as\nS\u000b\u000b\nL(!) =Z1\n\u00001dt\n2\u0019ei!thS\u000b\ni(t)S\u000b\ni(0)i=1\nNX\ncqeS\u000b\u000b\nc(q;!);\n(9)\nand are likewise isotropic in the D= 0 KLHM, i.e.\nS\u000b\u000b\nL(!) =SL(!). Their value at !\u00190 is experimen-\ntally highly relevant, as it is directly proportional to the\nexperimental NMR spin-lattice relaxation rate 1 =T1, pro-\nvided that hyper\fne form factors do not play an essential\nrole, as discussed in detail in Sec. V B. As mentioned pre-\nviously, we omit the singular q= 0,c= 0 spin di\u000busion\ncontribution, discussed further in Sec. V B.\nIn Fig. 7 we show the temperature evolution of the\nLSF over a broad range of T= 0:1{2:0 on theN= 30\nlattice. Apart from a pronounced low-energy peak arising\nfrom the dominant q= 0,c=\u00061 chiral \ructuations at\nT < 0:2 the LSF are quite temperature independent for\nT > 0:2, even at the relevant !\u00190 energy scale of\nNMR experiments. At T < 0:2, a drop of SL(!= 0) is\nobserved, which is again a signature of a \fnite spin triplet\ngap \u0001t>0, at least on \fnite-sized lattices [13{15].\nIt is instructive to compare the calculated LSF SL(!=\n0) to Moriya's Gaussian approximation [60] frequently\nused at high T\u001d1, but also extended to lower Tvia\nhigher-order corrections in the case of the D= 0 KLHM\n[29]. In a uniform Heisenberg spin-1 =2 model the LSF\nfrequency moments \u0016k=R\nd! !kSL(!), are exactly\nknown atT!1 , with the LSF sum rule \u00160= 1=4 and\u00162=z=8, wherez= 4 is the number of nearest-neighbors\nin the KL. These yield the expected != 0 value of the\nKLHM LSF under Gaussian line shape approximation\nSMoriya\nL (0) =\u00160p8\u0019\u00162\u00190:071; (10)\nwhich is reasonably close to the actual KLHM value\nSL(0)\u00190:055 atT= 2 calculated with the FTLM. We\nnote, though, that the frequency-dependent LSF SL(!)\nare not, in fact, Gaussian in shape, as is obvious from\nFig. 7, and T= 2 is not yet\u001d1.\nIV. DZYALOSHINSKII-MORIYA\nINTERACTIONS\nIn this section we consider an extension of the KLHM\nwith out-of-plane DM interactions 0 \u0014D\u00140:25 [Eq. (1)]\n(note that the dynamical response is not sensitive to the\nsign ofD) [25], which are relevant in many KL materials\n[26, 48, 50, 51, 54]. The out-of-plane Dleads to a uniax-\nially anisotropic chiral DSF eS\u000b\u000b\nc(q;!) with equal \u000b=x\nand\u000b=ycomponents that di\u000ber from the \u000b=zcom-\nponent, which has to be calculated separately. The same\nalso holds for all derived quantities, including the stan-\ndard DSFS\u000b\u000b(q;!) [Eq. (2)] and quantities in Eqs. (7).\nWe note that chiral spin operators eS\u000b\ncqwith\u000b=x;y\nare o\u000b-diagonal in the Sz\ntotbasis, which substantially\nincreases the overall computational complexity and re-\nquirements of FTLM compared to the \u000b=zcase, where\nthey are diagonal in the subspace. In particular, the em-\nployed reduced summation over Sz\ntotsubspaces (having\nlesser e\u000bect on diagonal correlations) appears to in\ruence\nmore the calculation of o\u000b-diagonal \u000b=x;ycorrelations.\nTo reduce di\u000berences we normalize \u000b=x;yresults by a\nscaling factor of 1 :15 to reproduce the \u000b=zsum rules\natD= 0.\nA. Dynamical spin structure factor\nIt is known that at low temperatures KLHM systems\ncan be signi\fcantly a\u000bected by the presence of additional\nDM interactions, with a quantum phase transition from\na SL GS to a 120 °AFM LRO GS with nonzero vector\nspin chirality when D >Dc\u00190:1 [52, 53]. This mainly\ncorresponds to a gradual softening of the dominant q= 0,\nc=\u00061 chiral \ructuations as Dincreases towards the\nquantum critical point Dc, beyond which these emerge\nas in-plane chiral 120 °AFM LRO.\nIn Fig. 8 we present the dominant q= 0,c=\u00061 chi-\nral DSF at a temperature T= 0:3 high enough to avoid\nlonger-ranged AFM correlations leading to strong \fnite-\nsize e\u000bects in our FTLM calculations. A \fnite D > 0\nsubstantially decreases the \u000b=zcomponent of the chiral\nDSF at low !, consistent with an increase of the e\u000bec-\ntive out-of-plane spin triplet gap \u0001z\nt. At the same time,\nthe\u000b=zspectra become sharper (more coherent) for7\n0.00.10.20.30.40.5-\n10 1 2 3 4 0.00.20.40.60.81.00.00.10.20.00.5(a) \nSzz±\n1(q = 0, ω) \nD = 0 \nD = 0.05 \nD = 0.1 \nD = 0.15 \nD = 0.2 \nD = 0.25 \nSxx±\n1(q = 0, ω)ω\n(b) \nωzzm\naxD\nModelH yperbola (User)E\nquation( k*dx) * sqrt(1 + ((x-x0)/dx)^2)P\nloto mega_maxx\n00 ± 0d\nx0 .05664 ± 0.0026k\n3 .04742 ± 0.02872R\neduced Chi-Sqr8 .28055E-5R\n-Square (COD)0 .99883A\ndj. R-Square0 .99853\nFigure 8. Chiral DSF's (a) eSzz\n\u00061(q= 0;!) and (b) eSxx\n\u00061(q=\n0;!) at a \fxed T= 0:3 and di\u000berent D= 0{0:25 on theN=\n30 lattice. Inset in (a) shows the frequency of the lower-energy\nmaximum of eSzz\n\u00061(q= 0;!) (symbols) with curves serving as\nguides to the eye.\nD > 0:1, i.e. beyond the quantum critical point, with\nthe energy of the spectral peak scaling nearly linearly as\n!max\u00193:0D[see inset in Fig. 8(a)]. This is consistent\nwith the linear scaling of the lower speci\fc heat peak\nTmax\u00190:91Dfound via the FTLM in Ref. [26], below\nwhich the spin correlation length \u0018increases substan-\ntially. The \u000b=x;ycomponents of the chiral DSF show\nthe latter e\u000bect quite clearly [Fig. 8(b)] as low-energy\noscillations due to \fnite-size magnon-like excitations be-\ncome visible at D&0:1 and increase in prominence as\nDincreases further. This indicates a considerable in-\ncrease in the spin correlation length \u0018 > 1 already at\nT&Tmaxwith increasing D>Dc. Concomitantly, there\nis a substantial increase of low- !intensity in \u000b=x;y\ncomponents of the chiral DSF, in contrast to a decrease\nin the\u000b=zcomponent, consistent with a softening of\nchiral \ructuations above a 120 °AFM GS with in-plane\nLRO spins [26, 48, 50, 52{54].\nB. Equal-time correlations and local spin\n\ructuations\nSimilar conclusions can be drawn from the tempera-\nture dependence of the chiral equal-time correlation func-\ntioneS\u000b\u000b\nc(q), which is de\fned by replacing the standard\nDSFS\u000b\u000b(q;!) in Eqs. (7) by the chiral DSF eS\u000b\u000b\nc(q;!).\nWe focus on the dominant q= 0,c=\u00061 correla-\ntions, which are shown in Fig. 9. We see that they\nare weakly T-dependent over the whole T > T fsrange\n0.30.40.50\n.00 .51 .01 .50.30.40.5 D = 0 \nD = 0.05 \nD = 0.1 \nD = 0.15 \nD = 0.2 \nD = 0.25(a)Szz±\n1(q = 0)T\n(b)Sxx±\n1(q = 0)Figure 9. The temperature dependence of chiral equal-time\nspin correlation functions (a) eSzz\n\u00061(q= 0) and (b) eSxx\n\u00061(q= 0)\nfor di\u000berent D= 0{0:25 on theN= 27 lattice.\nwhenD < Dc. The behavior changes qualitatively for\nD > Dc. While the \u000b=zcomponent remains rela-\ntively una\u000bected [Fig. 9(a)], the \u000b=x;ycomponents\nshow a strong increase below T.2D[Fig. 9(b)] consis-\ntent with the gradual onset of longer-range correlations\naroundT\u0018Tmax[26] and ultimate chiral AFM LRO at\nT= 0.\nFinally, we consider the temperature dependence of\nthe!= 0 LSF, which are directly relevant for NMR\nspin-lattice relaxation rate (1 =T1) experiments that we\ndiscuss in Sec. V B. Here we \fnd it useful to separately\nconsider the individual chiral LSF contributions\neS\u000b\u000b\nLc(!) =3\nNX\nqeS\u000b\u000b\nc(q;!); (11)\nto the full LSF S\u000b\u000b\nL(!) = (1=3)P\nceS\u000b\u000b\nLc(!). Note that\nthec= +1 andc=\u00001 chiral LSF are equal.\nIn Fig. 10 we present the calculated temperature de-\npendence of the != 0 chiral LSF for a range of D= 0{\n0:25 on theN= 30 lattice. We \fnd that the c=\u00061 chi-\nral LSF are highly sensitive to D, especially at T.2D\nwhere the\u000b=zcomponent is suppressed [Fig. 10(a)]\ndue to a shift of spectral intensity to higher !\u0018!max\n[Fig. 8(a)], while the \u000b=x;ycomponents are strongly\nenhanced [Fig. 10(c)] due to the gradual onset of longer-\nrange correlations at T\u0018Tmax[Fig. 8(b)]. On the other\nhand, components of the c= 0 chiral LSF are nearly\nequal and mostly insensitive to D, showing just a steady\nincrease with increasing temperature due to increasingly\nincoherent spin dynamics at T&1 [Fig. 10(b,d)].8\n0.000.020.040.06\n0.0 0.5 1.0 1.50.000.020.040.060.080.000.020.040.06\n0.0 0.5 1.0 1.50.000.020.040.06\n(a)\n D = 0\n D = 0.05\n D = 0.1\n D = 0.15\n D = 0.2\n D = 0.25Szz\nL ±1(ω = 0)\n(c)Sxx\nL ±1(ω = 0)\nT(b)\nSzz\nL0(ω = 0)\n(d)\nSxx\nL0(ω = 0)\nT\nFigure 10. The temperature dependence of d.c. chiral LSF (a) eSzz\nL\u00061(!= 0), (b) eSzz\nL0(!= 0), (c) eSxx\nL\u00061(!= 0), and (d)\neSxx\nL0(!= 0) for di\u000berent D= 0{0:25 on theN= 30 lattice.\nV. COMPARISON WITH EXPERIMENT\nIn this section we reinstate J6= 1 and SI units.\nA. Inelastic neutron scattering\nINS is a very powerful experimental technique as it\ndirectly probes the magnetic DSF S\u000b\u000b(q;!), with typi-\ncal interaction energies in KLHM materials J\u0018kB(60{\n230 K) = 5{20 meV in a convenient energy range for this\ntechnique. Unfortunately, most KL materials are not\nyet available in single-crystal form, therefore the intrinsic\nDSF anisotropy and q-dependence is often averaged out\nin experiment. To avoid this issue we concentrate on INS\nresults on single-crystal herbertsmithite [37], a material\nthat remains in a SL state down to the lowest measured\nkBT\u001910\u00004J. As the experimentally determined D=\n(0:04{0:08)J[50, 51] plays only a modest role against a\nmuch stronger J=kB\u0019190 K (16:4 meV) in equal-time\nproperties relevant for INS (Fig. 9), we compare INS re-\nsults with model calculations for D= 0. Moreover, low-\n!results may be strongly in\ruenced by structural and\nchemical disorder, especially at low kBT\u001cJ. We there-\nfore restrict ourselves to INS energies ~!>1 meV, above\nthe energy scale of impurity contributions, which mostly\ncontribute to a low- !quasielastic INS peak [37, 61].\nFirstly, we note that experimental frequency-\ndependent INS spectra show a broad maximum at ~!\u0018\n6 meV [37], which is consistent with the calculated low-\nenergy peak at ~!\u00190:3J= 5 meV (Figs. 3 and 4).\nSecondly, our calculations also nicely reproduce the q-dependence of the DSF integrated over a broad frequency\nwindow 1 meV <~!<11 meV (i.e. 0 :06J <~!<0:67J)\nalong the (\u00002;1 +K;0) cut in q-space, which shows the\nmost pronounced variation (Fig. 11). The position of the\nexperimental INS peak at ( \u00002;1;0) corresponds to the\nM point of the extended BZ and is well accounted for by\nour FTLM results. Having separated chiral contributions\n-1.0- 0.50 .00 .51 .00.00.10.20.3 \nHan et al. (2012) \nDimer model \nFTLM \nIntegrated S(q, ω) (arb. units)(\n-2, 1 + K, 0)/s49/s8211/s49/s49/s32/s109/s101/s86\nFigure 11. Low- Therbertsmithite INS measurements of the\nmagnetic DSF S(q;!) (symbols) integrated over 1 meV <\n~! < 11 meV along the ( \u00002;1 +K;0) cut in q-space from\nRef. [37]. The presented magnetic DSF was obtained from\nraw experimental data by dividing INS intensities by the free-\nCu2+magnetic form factor jF(q)j2[37, 62]. The experimental\nmagnetic DSF agrees well with q-interpolated FTLM calcu-\nlations atT= 0:1 on theN= 30 lattice (blue line), but\nsigni\fcantly worse with a toy model of independent singlet\ndimers [37] (dashed line).9\nat di\u000berent wave vectors, we can attribute this peak to\nthe dominant low-energy q= 0,c=\u00061 chiral \ructua-\ntions [Fig. 3 and Eq. (5)]. The position of the peak is\nalso consistent with expectations from the q-dependent\nd.c. susceptibility \u001f0(q), which is also sensitive mainly to\nlow-energy \ructuations, and which also has a very pro-\nnounced peak at the same wavevector (Fig. 6 and discus-\nsion in Sec. III B). Finally, we stress that not only the\nposition but also the width of the experimental INS peak\nis well reproduced by model calculations (Fig. 11), and is\nconsiderably smaller than the width predicted by a sim-\nple independent singlet dimer model [37]. This indicates\nthat the chiral AFM \ructuations in the KLHM have a\nnontrivial low- Tcorrelation length \u0018 > 1 that extends\nbeyond nearest KL neighbors.\nB. NMR spin-lattice relaxation rate\n1. Theory\nNMR spin relaxation experiments probe low-energy\nelectron spin \ructuations via the hyper\fne coupling be-\ntween nuclear and electron spins. In a crystal, the spin-\nlattice relaxation rate of a given nucleus is given by\n[63, 64]\n1\nT1=\r2\nn\n2Z1\n\u00001dt ei!0tX\nij\u000b\f(\u000e\u000b\f\u0000^B\u000b^B\f)h\u000eb\u000b\ni(t)\u000eb\f\nj(0)i;\n(12)\nwhere\rnis the nuclear gyromagnetic ratio, !0=\rnB\u001c\nJ=~is the nuclear Larmor angular frequency in an exter-\nnal \feld B,^B=B=jBjis a unit vector pointing along\nB, and\u000eb\u000b\ni=b\u000b\ni\u0000hb\u000b\nii=\u0000P\n\u0016A\u000b\u0016\niS\u0016\niis the e\u000bective\n\ructuating local \feld at the position of the nucleus due\nto hyper\fne coupling with the electron spin S\u0016\nivia the\nspeci\fc hyper\fne coupling tensor A\u000b\u0016\ni. De\fning the chi-\nral hyper\fne coupling tensor in q-space in analogy with\n[Eq. (3)] as\neA\u000b\u0016\ncq=1p\nNX\nneiq\u0001eRnh\nA\u000b\u0016\n(n;0)+\u0010cA\u000b\u0016\n(n;1)+\u0010\u0000cA\u000b\u0016\n(n;2)i\n;\n(13)\nwe can further succinctly express the NMR spin-lattice\nrelaxation rate in terms of the chiral DSF eS\u000b\f\nc(q;!)\n[Eq. (4)] as\n1\nT1=\u0019\r2\nnX\ncqtrn\neAy\ncq\u0001P?\u0001eAcq\u0001eSc(q;!0)o\n;(14)\nwhere the tensor P?= I\u0000^B\n^Bprojects onto a plane\northogonal to B, whileeAcqandeSc(q;!0) are 3\u00023 tensors\nwith components eA\u000b\u0016\ncqandeS\u000b\f\nc(q;!0), respectively.\nIn the simplest, yet experimentally highly relevant,\ncase of a nucleus coupled to z1spins of the KL triangle\nwith hyper\fne eigenaxes along the crystallographic axes\n[i.e. forA\u000b\u0016\n(n;k)=A\u000b\u000e\u000b\u0016\u000en;n0\u000ek\u0014z1], this further simpli\festo an expression involving only the chiral LSF eS\u000b\u000b\nLc(!0)\n[Eq. (11) and Fig. 10]\n1\nT1=\u0019\r2\nnX\n\u000bA2\n\u000b(1\u0000^B2\n\u000b)1\neT\u000b\u000b\n1;\n1\neT\u000b\u000b\n1=1\n31X\nc=\u00001fceS\u000b\u000b\nLc(!0);(15)\nwhere 1=eT\u000b\u000b\n1are directional contributions to the spin-\nlattice relaxation rate 1 =T1that depend on the number\nof spinsz1the nucleus is coupled to via the chiral form\nfactorsfcsummarized in Table I.\nIn Figs. 12(a{c) we show the impact of di\u000berent z1on\n1=eT\u000b\u000b\n1in more detail. We consider the \u000b=x;ycompo-\nnent due to spin \ructuations within the kagome plane,\nthe\u000b=zcomponent due to out-of-plane spin \ructua-\ntions, and a powder average of both, for both zero and\nlargeD= 0:25Jon theN= 30 lattice. We present our\nresults normalized to Moriya's Gaussian approximation\n[60] for the KLHM where eS\u000b\u000b;Moriya\nLc (0) =SMoriya\nL (0) for\nallcand\u000b[Eq. (10)], yielding\n1\neTMoriya\n1=~z1\n8p\u0019J: (16)\nFirstly, in the z1= 1 case [Fig. 12(a)], where each nucleus\nis coupled to a single magnetic ion, we have f0=f\u00061= 1\n(Table I), i.e. all chiralities contribute equally, and\n1=eT\u000b\u000b\n1=S\u000b\u000b\nL(!0\u00190) [Eq. (9)]. In the intermediate case\nofz1= 2 [e.g. when each nucleus is coupled equally to\ntwo spins on an exchange bond; see inset in Fig. 12(a)] we\nhavef0>f\u00061>0, where again all chiralities contribute\nto 1=eT\u000b\u000b\n1but chiral the c=\u00061 contributions are sup-\npressed compared to the c= 0 contribution [Fig. 12(b)].\nFinally, in the case of z1= 3 [e.g. when each nucleus\nis positioned symmetrically at or above the center of a\nKL triangle; see inset in Fig. 12(a)] we have f\u00061= 0, so\nthat thec=\u00061 \ructuations are completely \fltered out,\nand just the c= 0 chiral LSF [Figs. 10(b,d)] contribute,\nresulting in a nearly isotropic 1 =eT\u000b\u000b\n1= 3eS\u000b\u000b\nL0(!0\u00190)\nsteadily increasing with T[Fig. 12(c)].\nTable I. NMR chiral form factors fcin Eqs. (15) for coupling\nto di\u000berent numbers of spins z1in a single KL triangle and\nexamples of relevant nuclei in herbertsmithite, YCu 3(OH) 6Cl3\nand other compounds [see inset in Fig. 12(a)]. Note thatP\ncfc= 3z1.\nz1Nuclear position Nucleus Non-chiral f0Chiralf\u00061\n1 Magnetic ion (on-site)63,65Cu 1 1\n2 Exchange bond (NN)17O,1H 4 1\n3 Center of spin triangle35Cl 9 010\n0.0 0.5 1.0 1.50.00.20.40.60.80.00.20.40.60.8\n0.00.20.40.60.8\n0.0 0.5 1.0 1.50.00.20.40.60.8\n(d)35Cl in YCu3(OH)6Cl3 (×0.40)\n 4.7 T // c\n 4.7 T // a\n17O in herbertsmithite (×0.57)\n 9 T // c\n 9 T // a*\n 3.2 T // a* \nkBT/J\nTMoriya\n1/T1D = 0.25J\n Powder\n zz\n xx(a) z1 = 1TMoriya\n1/Tαα\n1\nD = 0\n(b) z1 = 2\nTMoriya\n1/Tαα\n1\nD = 0.25J\n Powder\n zz\n xxD = 0\n(c) z1 = 3TMoriya\n1/Tαα\n1\nkBT/JD = 0\nD = 0.25J\n Powder\n zz\n xxO2–\nCu2+\nCl–\nH+\nFigure 12. Directional spin-lattice relaxation rate contributions 1 =eT\u000b\u000b\n1[Eqs. (15) and Table I] normalized by Moriya's Gaussian\napproximation 1 =eTMoriya\n1 [Eq. (16)] for nuclei coupled to z1electron spins where (a) z1= 1 (63,65Cu-type), (b) z1= 2 (17O-type),\nand (c)z1= 3 (35Cl-type nuclei) calculated for D= 0 andD= 0:25Jon theN= 30 lattice. Shown are the \u000b=x;ycomponent\n(xx), the\u000b=zcomponent ( zz), and a powder average of both given by 1 =eTpowder\n1 = (1=3)P\n\u000b1=eT\u000b\u000b\n1. Representative nuclei\nin herbertsmithite and similar materials are shown on the inset in panel (a). (d) Symbols show the17O NMR spin-lattice\nrelaxation rate 1 =T1of herbertsmithite with J=kB= 190 K and z1= 2 from Ref. [38] (green) and the35Cl NMR spin-lattice\nrelaxation rate of YCu 3(OH) 6Cl3withJ=kB= 82 K [26] and z1= 3 from Ref. [56] (red), both normalized by Eq. (17). These\nvalues are further uniformly rescaled by a factor 0 :57 in the case of herbertsmithite and 0 :40 in the case of YCu 3(OH) 6Cl3.\nThese are compared to FTLM results at D= 0 andD= 0:25Jon theN= 30 lattice, with the curves taking appropriate\naverages over relevant directions \u000band chiral form factors fcin Eqs. (15) and Eq. (17).\n2. Experiment\nFirst we compare our FTLM model results with NMR\nexperiments on herbertsmithite, ZnCu 3(OH) 6Cl2. Even\nthough several experimental NMR spin relaxation stud-\nies have been carried out on this material over the years,\nsingle-crystal studies at kBT >0:1Jrelevant for compar-\nison with our model calculations are rare. In Fig. 12(d)\nwe summarize the17O NMR 1=T1results from Ref. [38]\nmeasured in the direction orthogonal to the kagome\nplanes (c-axis) and within the kagome planes ( a\u0003-axis).\nThe appropriate components of the hyper\fne coupling\ntensors (Aa;Aa\u0003;Ac) = (3:3ga;4:3ga;3:6gc) T are taken\nfrom Ref. [29]. Here ga= 2:14 andgc= 2:25 are in-\nplane and out-of-plane components, respectively, of the\nCu2+g-factor tensor at high- T[49]. As the oxygen nuclei\nare positioned symmetrically with respect to two neigh-\nboring magnetic Cu2+ions [inset in Fig. 12(a)], we have\nz1= 2 and the corresponding chiral form factors are\nf0= 4 andf\u00061= 1 (Table I). To compare our calcula-\ntions with experiment, we normalize all 1 =T1values toGaussian approximation [60]\n1\nTMoriya\n1=p\u0019\r2\nn~z1\n8JX\n\u000bA2\n\u000b(1\u0000^B2\n\u000b); (17)\nwhich can be obtained by inserting the directional\n1=eTMoriya\n1 [Eq. (16)] into the full 1 =T1[Eqs. (15)]. Like\nin the INS analysis in Sec. V A we compare experimental\nresults with FTLM calculations for D= 0, as the e\u000bect\nof the DM interaction on the chiral LSF at the experi-\nmentally determined value of D= (0:04{0:08)J[50, 51]\nis very small for all directions and chiralities (see the\nD= 0:05 curves in Fig. 10). The experimental 1 =T1\nalong the two crystallographic directions indeed coincide\nwhen normalized by Eq. (17), and their graduate decrease\nwith lowering Tnicely follows the theoretical prediction\ndown tokBT\u00190:3J. The downturn of the experimental\n17O NMR 1=T1below this temperature, which ultimately\nleads to 1=T1/T0:84belowkBT\u00190:05J[39], also seems\nto be qualitatively supported by our model calculations,\nwhere in the latter the downturn is the signature of a\nquite robust triplet gap \u0001 t>0 in the considered D= 0\nmodel system.\nThe second experimental example is the novel KL ma-11\nterial YCu 3(OH) 6Cl3, which, like herbertsmithite, has a\nnearest-neighbor Heisenberg exchange coupling J=kB=\n82 K that is by far the dominant isotropic magnetic inter-\naction [26]. However, unlike herbertsmithite, this mate-\nrial enters a chiral 120 °AFM LRO GS at kBTN= 0:15J\n[45, 47, 48], which is attributed to a sizable out-of-plane\nDM interaction D= 0:25J[26].35Cl NMR 1=T1results\nfrom Ref. [56] measured in the direction orthogonal to the\nkagome planes ( c-axis) and within the kagome planes ( a-\naxis), on one of the two chlorine crystallographic sites,\nare shown in Fig. 12(d). The chosen35Cl site is cou-\npled symmetrically with all three Cu2+spins on a given\nKL spin triangle [inset in Fig. 12(a)], similar to chlorine\nsites in herbertsmithite. The appropriate components of\nthe hyper\fne coupling tensor to a single electron spin are\nAa=Ac= 0:28 T. As evident by model calculations for a\nnucleus in such a symmetric z1= 3 position [Fig. 12(c)],\nthe anisotropy of the measured 1 =T1is minimal, sug-\ngesting that highly anisotropic chiral c=\u00061 local spin\n\ructuations [Figs. 10(a,c)] are indeed highly suppressed\nat the35Cl site, broadly consistent with the expected\nchiral form factors (Table I). Nevertheless, even though\nthe theoretically predicted trend of decreasing 1 =T1with\nloweringTis followed by experiment, the experimentally-\nobserved decrease is less pronounced [Fig. 12(d)]. As the\nc= 0 chiral LSF, which should represent the only contri-\nbution to 1=T1according to Table I, is expected to nearly\nvanish at low T[Figs. 10(b,d)], while the experimental\n1=T1does not, this suggests that a remnant c=\u00061 con-\ntribution must still a\u000bect the experimental 1 =T1to a cer-\ntain extent. This could be a telltale sign of reduced local\nthreefold rotational symmetry in YCu 3(OH) 6Cl3, similar\nto the recently discovered symmetry reduction in her-\nbertsmithite [49].\nFinally, we note that the experimental 1 =T1results\non both herbertsmithite and YCu 3(OH) 6Cl3need to be\nrescaled by factors of 0 :57 and 0:40, respectively, to\nachieve a quantitative match with model calculations\n[Fig. 12(d)]. This might be partly attributed to uncer-\ntainty in experimental parameters such as the hyper\fne\ncoupling constants. A further source of uncertainty is\nalso the spin di\u000busion contribution, i.e. the contribution\nfrom theq\u00190,c= 0 spin \ructuations, which we omit as\nmentioned in Sec. III A. In generic 2D systems at !0!0\nthis contribution might even be singular [59]. Never-\ntheless, in systems with strong AFM \ructuations, like\ncuprates [65] and KLHM materials, the spin di\u000busion\ncontribution to 1 =T1is generally considered to be rela-\ntively small and sizable only at high T. Still, it might\ncontribute a relevant quantitative correction to the cal-\nculated 1=T1.\nVI. SUMMARY AND OUTLOOK\nOur comprehensive numerical study of the dynamical\nspin correlations of the KL AFM via FTLM calculations\nhas led to several pertinent \fndings. By separating thechiral correlations ( c=\u00061) from non-chiral ones ( c= 0),\nwe have shown that former dominate the low-energy dy-\nnamics even of the isotropic D= 0 KLHM (Figs. 2\nand 3). These corresponds to \ructuations of the uni-\nform (q= 0) 120 °AFM order parameter, leading to a\npronounced low-energy response in the DSF S(q;!) at\nthe M point of the extended BZ. The dominant chiral\nDSF features a nontrivial frequency dependence charac-\nterized by a double-maximum structure that persists up\ntokBT\u0019J(Fig. 4), even though the lower-energy peak\ncorresponds to energies of only around 0 :3J. As a di-\nrect consequence of this low-energy peak, the d.c. sus-\nceptibility\u001f0(q) exhibits a pronounced peak at the M\npoint at low T(Fig. 6). In clear contrast, the equal-time\nS(q), which sums over all energies, exhibit a pronounced\nregion of high intensity that is spread out around the\nwhole extended BZ boundary, remains stable up to high\nkBT\u00182J, and has apparent weak maxima in the corner\nK points of the extended BZ (Fig. 5).\nAllowing for \fnite DM interactions perpendicular to\nthe kagome plane makes the chiral DSF anisotropic\n(Fig. 8). Such magnetic anisotropy mainly a\u000bects the\nq= 0,c=\u00061 chiral 120 °AFM \ructuations, which\nsoften at the quantum critical point Dc\u00190:1J. The\ncorresponding out-of-plane chiral DSF response spec-\ntraeSzz\n\u00061(q= 0;!) become more coherent with increas-\ningDwith an increase in an e\u000bective out-of-plane spin\ntriplet gap [Fig. 8(a)], while in-plane chiral DSF spectra\neSxx\n\u00061(q= 0;!) show enhanced low-energy \ructuations and\nlonger-range correlations [Figs. 8(b) and 9]. The change\nin local (i.e. integrated over q) spin \ructuations, which\nare highly relevant for local-probe experiments like NMR,\nfrom the isotropic D= 0 case is also dominated by chiral\nc=\u00061 \ructuations (Fig. 10).\nAll of the observed characteristic features of the KL\nantiferromagnet DSF can also be probed experimen-\ntally via INS and NMR spin-lattice relaxation measure-\nments. We critically compare our results to two most\nrelevant examples of the nearest-neighbor KL materials,\nthe archetypal herbertsmithite and the novel KL mate-\nrial YCu 3(OH) 6Cl3. The former possesses rather small\nDM magnetic anisotropy and lacks LRO down to the\nlowest experimentally accessible temperatures, while the\nlatter is characterized by a much larger DM anisotropy\nand chiral 120 °LRO at low T. Single-crystal KL INS\nmeasurements with the required q-space resolution are\nso far only available for herbertsmithite. These measure-\nments indeed show a broad low-energy peak [37] at en-\nergies that are entirely consistent with the lower-energy,\n0:3Jpeak that we \fnd numerically. Furthermore, our\nmodel calculations also convincingly reproduce the vari-\nation ofS(q;!) measured along the ( \u00002;1 +K;0)q-cut\nin Ref. [37] (Fig. 11). We \fnd that the peak is consider-\nably narrower than predicted by a simple singlet-dimer\ntoy model [37], which indicates that chiral AFM \ructua-\ntions in the KLHM have a \fnite low- Tcorrelation length\n\u0018>1.\nFurthermore,17O NMR spin-lattice relaxation rate12\n1=T1measurements on herbertsmithite [38] are reason-\nably reproduced by model calculations [Fig. 12(d)], show-\ning that the e\u000bect of small DM interactions that are\npresent in this compound on dynamical spin correla-\ntions is indeed small. The experimental T-dependence\nis well consistent with numerical result, in particular at\nkBT > 0:3J. The observed variation predominantly re-\n\rects the evolution of the non-chiral ( c= 0) \ructuations,\nas the contribution of the chiral ( c=\u00061) \ructuations is\npartly \fltered out on the symmetric position of the17O\nnuclei. The situation is even more extreme in the case\nof35Cl NMR spin-lattice relaxation rate 1 =T1measure-\nments on YCu 3(OH) 6Cl3[56], where chiral c=\u00061 \ructu-\nations should be completely \fltered out due to rotational\nsymmetry at the nuclear site. Indeed, we observe almost\nno anisotropy in experimental 1 =T1, however the experi-\nment notably deviates from theory at low T, suggesting\nthat the chiral contribution might still contribute, likely\ndue to reduced local rotational symmetry.\nOur study has demonstrated that detailed knowledge\nof the dynamical spin structure factor of KL AFM atT >0 can indeed provide invaluable insight into the na-\nture of its low-energy spin excitations and represent a\nlink to numerous theoretical studies of the ground state of\nthis enigmatic model. Especially intriguing is the robust,\nyet hitherto underappreciated, chiral nature of the dom-\ninant spin \ructuations. The scope of our results is fur-\nther extended by the inclusion of experimentally highly-\nrelevant DM interactions with nontrivial consequences.\nWe have furthermore demonstrated that our unbiased\nstate-of-the-art numerical calculations provide a reliable\nbasis upon which past and future experiments on kagome\nmaterials can be judged and interpreted.\nACKNOWLEDGMENTS\nWe acknowledge the \fnancial support of the Slove-\nnian Research Agency through programs P1-0044 and\nP1-0125, and projects N1-0088, Z1-1852, N1-0148, and\nJ1-2461.\n[1] P. A. Lee, An end to the drought of quantum spin liquids,\nScience 321, 1306 (2008).\n[2] L. Balents, Spin liquids in frustrated magnets, Nature\n464, 199 (2010).\n[3] L. Savary and L. Balents, Quantum spin liquids: A re-\nview, Rep. Prog. Phys. 80, 016502 (2017).\n[4] C. Broholm, R. J. Cava, S. A. Kivelson, D. G. Nocera,\nM. R. Norman, and T. Senthil, Quantum Spin Liquids,\nScience 367, 263 (2020).\n[5] F. Mila, Low-energy sector of the kagome antiferromag-\nnet, Phys. Rev. Lett. 81, 2356 (1998).\n[6] C. Waldtmann, H. U. Everts, B. Bernu, C. Lhuillier,\nP. Sindzingre, P. Lecheminant, and L. Pierre, First ex-\ncitations of the spin 1/2 Heisenberg antiferromagnet on\nthe kagom\u0013 e lattice, Eur. Phys. J. B 2, 501 (1998).\n[7] R. R. P. Singh and D. A. Huse, Ground state of the spin-\n1/2 kagome-lattice Heisenberg antiferromagnet, Phys.\nRev. B 76, 180407(R) (2007).\n[8] R. R. P. Singh and D. A. Huse, Triplet and singlet excita-\ntions in the valence bond crystal phase of the kagome lat-\ntice Heisenberg model, Phys. Rev. B 77, 144415 (2008).\n[9] S. Yan, D. A. Huse, and S. R. White, Spin-liquid ground\nstate of the S=1/2 kagome Heisenberg antiferromagnet,\nScience 322, 1173 (2011).\n[10] A. M. L auchli, J. Sudan, and E. S. S\u001crensen, Ground-\nstate energy and spin gap of spin-1/2 Kagom\u0013 e-Heisenberg\nantiferromagnetic clusters: Large-scale exact diagonal-\nization results, Phys. Rev. B 83, 212401 (2011).\n[11] Y. Iqbal, F. Becca, and D. Poilblanc, Valence-bond crys-\ntal in the extended kagome spin-1/2 quantum Heisenberg\nantiferromagnet: A variational Monte Carlo approach,\nPhys. Rev. B 83, 100404(R) (2011).\n[12] S. Depenbrock, I. P. McCulloch, and U. Schollw ock, Na-\nture of the Spin-Liquid Ground State of the S= 1=2\nHeisenberg Model on the Kagome Lattice, Phys. Rev.\nLett. 109, 067201 (2012).[13] J. Schnack, J. Schulenburg, and J. Richter, Magnetism of\ntheN= 42 kagome lattice antiferromagnet, Phys. Rev.\nB98, 094423 (2018).\n[14] A. M. L auchli, J. Sudan, and R. Moessner, S=1/2 kagome\nHeisenberg antiferromagnet revisited, Phys. Rev. B 100,\n155142 (2019).\n[15] P. Prelov\u0014 sek, K. Morita, T. Tohyama, and J. Herbrych,\nVanishing Wilson ratio as the hallmark of quantum spin-\nliquid models, Phys. Rev. Research 2, 023024 (2020).\n[16] Y. Ran, M. Hermele, P. A. Lee, and X.-G. Wen,\nProjected-Wave-Function Study of the Spin-1 =2 Heisen-\nberg Model on the Kagom\u0013 e Lattice, Phys. Rev. Lett. 98,\n117205 (2007).\n[17] Y. Iqbal, F. Becca, S. Sorella, and D. Poilblanc, Gap-\nless spin-liquid phase in the kagome spin-1\n2Heisenberg\nantiferromagnet, Phys. Rev. B 87, 060405(R) (2013).\n[18] Z. Y. Xie, J. Chen, J. F. Yu, X. Kong, B. Normand,\nand T. Xiang, Tensor renormalization of quantum many-\nbody systems using projected entangled simplex states,\nPhys. Rev. X 4, 011025 (2014).\n[19] Y. Iqbal, D. Poilblanc, and F. Becca, Vanishing spin\ngap in a competing spin-liquid phase in the kagome\nHeisenberg antiferromagnet, Phys. Rev. B 89, 020407(R)\n(2014).\n[20] Y.-C. He, M. P. Zaletel, M. Oshikawa, and F. Poll-\nmann, Signatures of Dirac Cones in a DMRG Study of\nthe Kagome Heisenberg Model, Phys. Rev. X 7, 031020\n(2017).\n[21] H. J. Liao, Z. Y. Xie, J. Chen, Z. Y. Liu, H. D. Xie, R. Z.\nHuang, B. Normand, and T. Xiang, Gapless spin-liquid\nground state in the s= 1=2 kagome antiferromagnet,\nPhys. Rev. Lett. 118, 137202 (2017).\n[22] G. Misguich and P. Sindzingre, Magnetic suscepti-\nbility and speci\fc heat of the spin-1/2 Heisenberg\nmodel on the kagome lattice and experimental data on\nZnCu 3(OH) 6Cl2, Eur. Phys. J. B 59, 305 (2007).13\n[23] B. Bernu, L. Pierre, K. Essa\f, and L. Messio, E\u000bect of\nperturbations on the kagome s=1\n2antiferromagnet at\nall temperatures, Phys. Rev. B 101, 140403 (2020).\n[24] M. Rigol and R. R. P. Singh, Magnetic Susceptibility\nof the Kagome Antiferromagnet ZnCu 3(OH)6Cl2, Phys.\nRev. Lett. 98, 207204 (2007).\n[25] M. Rigol and R. R. P. Singh, Kagome lattice antiferro-\nmagnets and Dzyaloshinsky-Moriya interactions, Phys.\nRev. B 76, 184403 (2007).\n[26] T. Arh, M. Gomil\u0014 sek, P. Prelov\u0014 sek, M. Pregelj,\nM. Klanj\u0014 sek, A. Ozarowski, S. J. Clark, T. Lancaster,\nW. Sun, J.-X. Mi, and A. Zorko, Origin of magnetic or-\ndering in a structurally perfect quantum kagome antifer-\nromagnet, Phys. Rev. Lett. 125, 027203 (2020).\n[27] T. Shimokawa and H. Kawamura, Finite-temperature\ncrossover phenomenon in the S= 1=2 antiferromagnetic\nHeisenberg model on the kagome lattice, J. Phys. Soc.\nJpn.85, 6 (2016).\n[28] N. E. Sherman and R. R. P. Singh, Structure factors of\nthe kagome-lattice Heisenberg antiferromagnets at \fnite\ntemperatures, Phys. Rev. B 97, 014423 (2018).\n[29] N. E. Sherman, T. Imai, and R. R. P. Singh, Nuclear\nrelaxation rates in the herbertsmithite kagome antifer-\nromagnets ZnCu 3(OH)6Cl2, Phys. Rev. B 94, 140415\n(2016).\n[30] Z. Hao and O. Tchernyshyov, Structure factor of low-\nenergy spin excitations in a S= 1=2 kagome antiferro-\nmagnet, Phys. Rev. B 81, 1 (2010).\n[31] M. Punk, D. Chowdhury, and S. Sachdev, Topological ex-\ncitations and the dynamic structure factor of spin liquids\non the kagome lattice, Nat. Phys. 10, 289 (2014).\n[32] W. Zhu, S.-s. Gong, and D. N. Sheng, Identifying spinon\nexcitations from dynamic structure factor of spin-1/2\nHeisenberg antiferromagnet on the Kagome lattice, Proc.\nNatl. Acad. Sci. 116, 5437 (2019).\n[33] T. Shimokawa, K. Watanabe, and H. Kawamura, Static\nand dynamical spin correlations of the S=1\n2random-\nbond antiferromagnetic Heisenberg model on the trian-\ngular and kagome lattices, Phys. Rev. B 92, 1 (2015).\n[34] M. P. Shores, E. A. Nytko, B. M. Bartlett, and D. G.\nNocera, A structurally perfect s = 1/2 kagom\u0013 e antiferro-\nmagnet, J. Am. Chem. Soc. 127, 13462 (2005).\n[35] P. Mendels and F. Bert, Quantum kagome antiferro-\nmagnet ZnCu 3(OH) 6Cl2, J. Phys. Soc. Jpn. 79, 011001\n(2010).\n[36] M. R. Norman, Colloquium: Herbertsmithite and the\nsearch for the quantum spin liquid, Rev. Mod. Phys. 88,\n041002 (2016).\n[37] T. H. Han, J. S. Helton, S. Chu, D. G. Nocera, J. A.\nRodriguez-Rivera, C. Broholm, and Y. S. Lee, Fraction-\nalized excitations in the spin-liquid state of a kagome-\nlattice antiferromagnet, Nature 492, 406 (2012).\n[38] M. Fu, T. Imai, T.-H. Han, and Y. S. Lee, Evidence for a\ngapped spin-liquid ground state in a kagome Heisenberg\nantiferromagnet, Science 350, 655 (2015).\n[39] P. Khuntia, M. Velazquez, Q. Barth\u0013 elemy, F. Bert,\nE. Kermarrec, A. Legros, B. Bernu, L. Messio, A. Zorko,\nand P. Mendels, Gapless ground state in the archetypal\nquantum kagome antiferromagnet ZnCu 3(OH) 6Cl2, Nat.\nPhys. 16, 469 (2020).\n[40] Z. Hiroi, M. Hanawa, N. Kobayashi, M. Nohara, H. Tak-\nagi, Y. Kato, and M. Takigawa, Spin-1/2 Kagome-Like\nLattice in Volborthite Cu 3V2O7(OH) 2·2H2O, J. Phys.\nSoc. Jpn. 70, 3377 (2001).[41] B. F\u0017 ak, E. Kermarrec, L. Messio, B. Bernu, C. Lhuil-\nlier, F. Bert, P. Mendels, B. Koteswararao, F. Bouquet,\nJ. Ollivier, A. D. Hillier, A. Amato, R. H. Colman, and\nA. S. Wills, Kapellasite: A kagome quantum spin liq-\nuid with competing interactions, Phys. Rev. Lett. 109,\n037208 (2012).\n[42] Y. Li, B. Pan, S. Li, W. Tong, L. L., Z. Yang, J. Wang,\nZ. Chen, Z. Wu, and Q. Zhang, Gapless quantum spin\nliquid in the S = 1/2 anisotropic kagome antiferromagnet\nZnCu 3(OH) 6SO4, New J. Phys. 16, 093011 (2014).\n[43] M. Gomil\u0014 sek, M. Klanj\u0014 sek, M. Pregelj, F. C. Coomer,\nH. Luetkens, O. Zaharko, T. Fennell, Y. Li, Q. M. Zhang,\nand A. Zorko, Instabilities of spin-liquid states in a quan-\ntum kagome antiferromagnet, Phys. Rev. B 93, 060405\n(2016).\n[44] Z. Feng, Z. Li, X. Meng, W. Yi, Y. Wei, J. Zhang, Y. C.\nWang, W. Jiang, Z. Liu, S. Li, F. Liu, J. Luo, S. Li, G. Q.\nZheng, Z. Y. Meng, J. W. Mei, and Y. Shi, Gapped Spin-\n1/2 Spinon Excitations in a New Kagome Quantum Spin\nLiquid Compound Cu 3Zn(OH) 6FBr, Chin. Phys. Lett.\n34(2017).\n[45] A. Zorko, M. Pregelj, M. Klanj\u0014 sek, M. Gomil\u0014 sek,\nZ. Jagli\u0014 ci\u0013 c, J. S. Lord, J. A. T. Verezhak, T. Shang,\nW. Sun, and J.-X. Mi, Coexistence of magnetic order\nand persistent spin dynamics in a quantum kagome anti-\nferromagnet with no intersite mixing, Phys. Rev. B 99,\n214441 (2019).\n[46] W. Sun, Y.-X. Huang, S. Nokhrin, Y. Pan, and J.-X.\nMi, Perfect Kagom\u0013 e lattices in YCu 3(OH) 6Cl3: a new\ncandidate for the quantum spin liquid state, J. Mater.\nChem. C 4, 8772 (2016).\n[47] Q. Barth\u0013 elemy, P. Puphal, K. M. Zoch, C. Krell-\nner, H. Luetkens, C. Baines, D. Sheptyakov, E. Ker-\nmarrec, P. Mendels, and F. Bert, Local study\nof the insulating quantum kagome antiferromagnets\nYCu 3(OH)6OxCl3\u0000x(x= 0;1=3), Phys. Rev. Materials\n3, 074401 (2019).\n[48] A. Zorko, M. Pregelj, M. Gomil\u0014 sek, M. Klanj\u0014 sek, O. Za-\nharko, W. Sun, and J.-X. Mi, Negative-vector-chirality\n120\u000espin structure in the defect- and distortion-free\nquantum kagome antiferromagnet YCu 3(OH)6Cl3, Phys.\nRev. B 100, 144420 (2019).\n[49] A. Zorko, M. Herak, M. Gomil\u0014 sek, J. van Tol,\nM. Vel\u0013 azquez, P. Khuntia, F. Bert, and P. Mendels,\nSymmetry Reduction in the Quantum Kagome Antifer-\nromagnet Herbertsmithite, Phys. Rev. Lett. 118, 017202\n(2017).\n[50] A. Zorko, S. Nellutla, J. van Tol, L. C. Brunel, F. Bert,\nF. Duc, J. C. Trombe, M. A. de Vries, A. Harrison,\nand P. Mendels, Dzyaloshinsky-Moriya anisotropy in the\nspin-1/2 kagome compound ZnCu 3(OH) 6Cl2, Phys. Rev.\nLett. 101, 026405 (2008).\n[51] S. El Shawish, O. C\u0013 epas, and S. Miyashita, Electron spin\nresonance in S= 1=2 antiferromagnets at high tempera-\nture, Phys. Rev. B 81, 224421 (2010).\n[52] O. C\u0013 epas, C. M. Fong, P. W. Leung, and C. Lhuillier,\nQuantum phase transition induced by Dzyaloshinskii-\nMoriya interactions in the kagome antiferromagnet,\nPhys. Rev. B 78, 140405(R) (2008).\n[53] M. Elhajal, B. Canals, and C. Lacroix, Symmetry break-\ning due to Dzyaloshinsky-Moriya interactions in the\nkagom\u0013 e lattice, Phys. Rev. B 66, 014422 (2002).\n[54] A. Zorko, F. Bert, A. Ozarowski, J. van Tol, D. Boldrin,\nA. S. Wills, and P. Mendels, Dzyaloshinsky-Moriya in-14\nteraction in vesignieite: A route to freezing in a quan-\ntum kagome antiferromagnet, Phys. Rev. B 88, 144419\n(2013).\n[55] J. Jakli\u0014 c and P. Prelov\u0014 sek, Finite-temperature properties\nof doped antiferromagnets, Adv. Phys. 49, 1 (2000).\n[56] T. Arh, M. Klanj\u0014 sek, M. Pregelj, M. Gomil\u0014 sek, and\nA. Zorko, Local insight into spin correlations in the\nkagome antiferromagnet YCu 3(OH) 6Cl3, manuscript in\npreparation.\n[57] J. Jakli\u0014 c and P. Prelov\u0014 sek, Lanczos method for the calcu-\nlation of \fnite-temperature quantities in correlated sys-\ntems, Phys. Rev. B 49, 5065 (1994).\n[58] P. Prelov\u0014 sek and J. Bon\u0014 ca, Ground State and Finite\nTemperature Lanczos Methods, in Strongly Correlated\nSystems|Numerical Methods , edited by A. Avella and\nF. Mancini (Springer, Berlin, 2013).\n[59] A. Sokol, E. Gagliano, and S. Bacci, Theory of nuclear\nspin{lattice relaxation in La 2CuO 4at high temperatures,\nPhys. Rev. B 47, 14646 (1993).\n[60] T. Moriya, Nuclear magnetic relaxation in antiferromag-\nnets, Prog. Theor. Phys. 16, 23 (1956).\n[61] T.-H. Han, M. R. Norman, J.-J. Wen, J. A. Rodriguez-Rivera, J. S. Helton, C. Broholm, and Y. S. Lee, Cor-\nrelated impurities and intrinsic spin-liquid physics in\nthe kagome material herbertsmithite, Phys. Rev. B 94,\n060409 (2016).\n[62] G. Shirane, S. Shapiro, and J. Tranquada, Neutron Scat-\ntering with a Triple-Axis Spectrometer: Basic Techniques\n(Cambridge University Press, Cambridge, 2006).\n[63] M. Horvati\u0013 c, Magnetic nuclear spin-lattice relaxation in\nNMR of orthorhombic crystals in the presence of strong\nquadrupole coupling, J. Phys. Condens. Matter. 4, 5811\n(1992).\n[64] M. Horvati\u0013 c and C. Berthier, NMR Studies of Low-\nDimensional Quantum Antiferromagnets, in High Mag-\nnetic Fields: Applications in Condensed Matter Physics\nand Spectroscopy , edited by C. Berthier, L. P. L\u0013 evy, and\nG. Martinez (Springer Berlin Heidelberg, Berlin, Heidel-\nberg, 2001) pp. 191{210.\n[65] T. Imai, C. P. Slichter, K. Yoshimura, and K. Kosuge,\nLow frequency spin dynamics in undoped and Sr-doped\nLa2CuO 4, Phys. Rev. Lett. 70, 1002 (1993)." }, { "title": "1505.00106v1.Spin_charge_transport_driven_by_magnetization_dynamics_on_disordered_surface_of_doped_topological_insulators.pdf", "content": "arXiv:1505.00106v1 [cond-mat.mes-hall] 1 May 2015Spin-charge transport driven by magnetization dynamics on\ndisordered surface of doped topological insulators\nK. Taguchi, K. Shintani, and Y. Tanaka\nDepartment of Applied Physics, Nagoya University, Nagoya, 464-8603, Japan\nand CREST, Japan Science and Technology Corporation (JST), Nagoya 464-8603, Japan\n(Dated: June 4, 2021)\nAbstract\nWe theoretically study the spin and charge generation along with the electron transport on a\ndisordered surface of a doped three-dimensional topologic al insulator/magnetic insulator junction\nby using Green’s function techniques. We find that the spin an d charge current are induced by not\nonly local but also by nonlocal magnetization dynamics thro ugh nonmagnetic impurity scattering\non the disordered surface of the doped topological insulato r. We also clarify that the spin current\nas well as charge density are induced by spatially inhomogen eous magnetization dynamics, and\nthe spin current diffusively propagates on the disordered sur face. Using these results, we discuss\nboth local and nonlocal spin torques before and after the spi n and spin current generation on the\nsurface, and provide a procedure to detect the spin current.\n1I. INTRODUCTION\nIn spintronics, the mutual control of the direction and the flow of spin is a central issue\nfor wide applications. The flow of spin, i.e., spin current, is the differen ce between the\ncurrents of up and down-spin conduction electrons. It is known th at the spin current is\ninduced in the setup of the ferromagnetic metal (FM)/normal met al (NM) junction1–3. Its\norigin is due to the magnetization dynamics in ferromagnet, which tra nsfers the spin angular\nmomentum of the magnetization into that of the conduction electro ns. The transfer of the\nspin angular momentum is called spin-pumping. Here, the spin-pumping of magnetization\ndynamics generates the spin-current in the NM, and the spin-curr ent can be converted into\ncharge current through spin-orbit interactions1–5.\nTopological insulator (TI) is a new class of materials which has a gaples s surface state,\ndubbed as the helical surface state, in which the spin and momentum are locked by the spin-\norbit interactions6–8. On the surface of the TI, the direction of charge current and th at of\nthe spin of conduction electrons can be mutually manipulated by an ap plied electromagnetic\nfield through the spin-momentum locking. There have been many the oretical works in\nhybrid systems including superconducting junctions on the surfac e of TI stimulated by the\nexotic surface state7,9–12. In the TI/FM junctions, the anomalous charge-spin transport12–15,\nthe anomalous tunnel conductance16–20, the giant magneto resistance21–23, and the current-\ninduced spin-transfer torque24,25have been studied up to now. The exotic phenomena are\ntriggered in the presence of static magnetization and an applied elec tromagnetic field. The\nmagnetizationoftheferromagnetplaystheroleofaneffectivevec torpotentialforconduction\nelectrons, which is like a vector potential of electromagnetic fields. Owing to the effective\nvector potential, the time-derivative of the magnetization can be r egarded as an effective\nelectric field, and the magnetization dynamics generates charge cu rrent on the surface of the\nTI/FM junction even in the absence of electromagnetic fields26–29. This is called as the spin-\ncharge conversion. The direction of the induced charge current is perfectly perpendicular\nto the magnetization dynamics due to the spin-momentum locking. Th e relation between\nthe direction of the magnetization and the induced charge current can be a characteristic\nproperty on the surface of the TI. The property of the spin-pum ping on the surface of TI\ncan be applicable for spintronics devices.\nExisting works of the spin-charge conversion have been done in the case of a clean sur-\n2face of the TI, namely the ballistic transport regime. However, the actual charge trans-\nport on the surface of the TI is in the diffusive regime due to the nonm agnetic impurity\nscattering21–25,30–32. Since Burkov et al., have predicted not only the local but also the no n-\nlocal current onthe disordered surface of the TI inthe presence of the applied electric field21;\nwe can naturally expect nonlocal current is driven by the magnetiza tion dynamics even on\nthe disordered surface of the TI in the presence of the magnetiza tion dynamics.\nIn this article, we study the charge-spin transport due to the mag netization dynamics on\nthe disordered surface of the three-dimensional doped TI/magn etic insulator (MI) junction,\nas shown in Fig. 1, where we show that charge current and spin polar ization on the surface\nof the TI are induced not only by a local, but also by a nonlocal magnet ization dynamics.\nBesides this, we clarify that the spin current is driven by the dynamic s of the spatially\ninhomogenous magnetization, and the spin current diffusively propa gates on the surface.\nThe magnitude of the spin current reflects the spatially inhomogeno us spin structure of the\nMI. The directions of the spin flow and the spin projection of the spin current are perfectly\nlinked by the spin-momentum locking on the surface of TIs. The pres ent features may serve\nas a guide to fabricate future spintronics devices based on the sur face of TIs with magnetic\nsubstance.\nThe merit of the choice of MI on the surface of the TI instead of met allic ferromagnet is\nto prevent the induced charge current going through the bulk of t he MI. Then, we can focus\non the charge transport on the surface of TI. Besides, the Gilber t damping constant in MIs\ntends to be smaller than that in ferromagnetic metals. The small valu e of the damping in\nMIs can be useful for the detection of the spin current on the diso rdered surface of the TI,\nas discussed in sec. VC\nII. MODEL\nWe consider conduction electrons coupled to an effective localized sp in on the disordered\nsurface of the three-dimensional doped TI attached with the MI, as shown in Fig. 1. The\nsetup in Fig. 1 is similar to a system, where conduction electrons coup le with the magnetic\nmoments of ferromagnetic metals deposited on the surface of the TI12,16,17. We expect\nthat on the surface of the TI, the effective localized spin ( S) can be produced from the\nmagnetization in the MI through magnetic proximity effects. In the f ollowing, we use the\n3Hamiltonian, describing the surface of the TI with MI, given by\nH=HTI+Hsd+Vimp, (1)\nwhere the first term in Eq. (1), HTIis Hamiltonian of the conduction electrons on one of\nthe surface of the doped TI without Sas\nHTI=/integraldisplay\ndxψ†[−i/planckover2pi1vF(ˆσ×∇)z−ǫF]ψ, (2)\nHere,ψ†≡ψ†(x,t) = (ψ†\n↑ψ†\n↓), andψare the creation and annihilation operators of the\nconduction electron, respectively (where indices ↑and↓represent spin), ǫFis the Fermi\nenergy, and vFis the Fermi velocity of the bare electron on the surface of the dop ed TI. The\nˆσis the Pauli matrices in spin space. The second term of Eq. (1), Hsd, shows the exchange\ninteraction between the conduction electron spin s=1\n2ψ†σψand the localized spin Son\nthe disordered surface of the doped TI, as described by\nHsd=−/integraldisplay\ndxJsdψ†S·ˆσψ, (3)\nwhereJsd>0 is the exchange coupling constant. The localized spin Scan be described\nby the magnetization of the MI as S=−(S/M)M, whereSandMare the magnitude of\nthe localized spin and of the magnetization, respectively. We conside r that in general, the\nlocalized spin S≡S(x,t) depends on the time and position on the surface of the TI. The\nS(x,t) changes slowly compared with the electron transport relaxation t ime (τ) and varies\nin space compared with the electron mean-free path ( ℓ). We expect that from the Eqs.\n(2)-(3), the in-plane component of the localized spin, S/bardbl≡S−Szz, can play the role of\nthe effective vector potential for the conduction electrons on th e surface. The out-of plane\ncomponent of the localized spin Szplays a role to open the energy gap of the dispersion on\nthe surface of the doped TI. We assume that the band gap opened bySzis smaller than the\nFermi energy on the surface of the doped TI, i.e.,ǫF−JsdSz>0. The third term of Eq. (1),\nVimp=Ni/summationdisplay\nj=1/integraldisplay\ndxUiψ†ψ, (4)\nrepresents nonmagnetic impurity scattering on the disordered su rface of the doped TI. The\nimpurity scattering causes the relaxation time τof the transport of conduction electrons on\nthesurfaceoftheTI. Here Ui=uiδ(x−rj)isadelta-functiontype potential, uiisapotential\n4FIG. 1: (Color Online) A setup of spin-charge current genera tion due to magnetization dynamics\nin MIs deposited on disordered surface of a doped TI.\nenergy density, rjis the position of impurities, and Nishows the number of impurities. The\ncontribution of Vican be treated by the impurity average.\nWe will calculate the spin current and charge current due to the spin -pumping in the\nlinear response to Sunder the condition Jsd≪/planckover2pi1/τ. We expect that the condition could\nbe realized on a metallic disordered surface of the TI, which satisfies /planckover2pi1/(ǫFτ)≪1. For\nexample, the exchange coupling Jsdcan be estimated by Jsd≃6meV28for Ni81Fe19/Bi2Te3\njunction. If/planckover2pi1\nǫFτ<10−2is satisfied, the perturbation can be accessible.\nA. Renormalization of the Fermi velocity\nGreen’s function on disordered surface of doped TI can be describ ed by using HTIin-\ncludingVimpwithin the self-consistent Born approximation of Vimpas\nˆgk,ω= [/planckover2pi1ω−{/planckover2pi1vF(ˆσ×k)z−ǫF}−ˆΣk,ω]−1, (5)\nwhereˆΣk,ωis the self-energy within the Born approximation given by\nˆΣk,ω=ni/summationdisplay\nk′|uk−k′|2ˆgk′,ω. (6)\nHere,ˆΣk,ωsatisfiestheWard-Takahashiidentity33. Toestimatethevalueof ˆΣk,ω, weconsider\nthek′-dependence of uk−k′, which plays the role to prevent the ultraviolet-divergence over\na large momentum in the k′-integration25,34. When ˆΣk,ωcan be described by 2 ×2 matrix\nˆΣk,ω= Σ0+ˆΣ/bardbl+ Σzˆσz,ˆΣ is estimated, where Σ 0and Σzare independent of k, while\n5Σ/bardbl≡Σxˆσx+Σyˆσydepends on k. Then, the Green’s function ˆ gk,ωis described by Σ within\nthe Born approximation as25,34\nˆgr\nk,ω=/bracketleftbigg\n/planckover2pi1ω−{/planckover2pi1˜vF(ˆσ×k)z−ǫF}+i/planckover2pi1\n2τ/bracketrightbigg−1\n, (7)\nwhere ˆgr\nk,ωrepresents the retarded Green’s function. From Eq. (7), the Fe rmi velocity is\nrenormalized by Σ/bardbl, and the renormalized Fermi velocity is represented by ˜ vF=vF/(1+ξ),\nwhereξ=niu2\n0/(4π/planckover2pi12v2\nF) is a small value depending on the relaxation time35. The last term\nin Eq. (7) is caused by the retarded component of Im[Σ 0].\nEquation (7) indicates the Green’s function on the disordered surf ace of the doped TI\nestimated within the Born approximation of Vimp. Therefore, we could expect that /planckover2pi1˜vF(ˆσ×\nk)z−ǫFin the Eq. (7) corresponds to the dispersion on the disordered sur face of the TI.\nThe dispersion is different from that of HTIwithoutVimp. In the following work, we will use\nan effective Hamiltonian ˜HTIobtained by replacing vFwith ˜vFin Eq. (2). This replacement\nis needed for satisfying the charge conservation law on the disorde red surface of the doped\nTI36.\nIII. SPIN CURRENT DUE TO MAGNETIZATION DYNAMICS\nIn this section, we show spin current driven by magnetization dynam ics on the disordered\nsurface of the doped TI/MI junction. Here, the spin current and charge density are mutually\nrelated each other, because of the spin-momentum locking on the s urface of the TI.\nA. Definition of spin current on the surface of topological insulators\nIn order to derive the spin current on the disordered surface of t he doped TI, we demon-\nstrate the definition of the spin current. The spin current jα\niis defined from\n∂tsα+∇ijα\ni=Tα, (8)\nwheresα=1\n2∝an}b∇acketle{tψ†ˆσαψ∝an}b∇acket∇i}htis the spin density, jα\nishows the spin current density, and Tαis the\nspin relaxation torque on the surface. Here the subscript and sup erscript ofjα\nirepresent the\ndirection of flow and spin of the spin current, respectively. From Eq s. (1)-(5),jα\niandTα\nare given by\njα\ni=˜vF\n2ǫzαi∝an}b∇acketle{tψ†ψ∝an}b∇acket∇i}ht=˜vF\n2eǫzαiρe. (9)\n6with the Levi-Civita symbol ǫzαi. In the above equation, we used the commutation relation:\n[ψ†,H] =−i/planckover2pi1˜vFǫzαℓ(∇ℓψ†)σα+Jsdψ†Sασα,\n[ψ,H] =−i/planckover2pi1˜vFǫzαℓσα(∇ℓψ)−JsdSασαψ.\nFrom Eq. (9), the spin current is proportional to the charge dens ityρe=e∝an}b∇acketle{tψ†ψ∝an}b∇acket∇i}ht, where\ne <0 is the charge of electrons. Moreover, the directions of the spin a nd flow of the spin\ncurrent are perpendicular to each other because of the spin-mom entum locking. The spin\nrelaxation torque is derived from Eqs. (8)-(9). The torque can be separated as\nTα=Tα\nTI+Tα\nsd, (10)\nwhereTα\nTIandTα\nsdare spin relaxation torque caused by HTIandHsd, respectively. Here, Tα\nTI\nandTα\nsdare given by\nTα\nTI=i˜vF\n2ǫβzℓǫβαν∝an}b∇acketle{t(∇ℓψ†)ˆσνψ−ψ†ˆσν∇ℓψ∝an}b∇acket∇i}ht, (11)\nTα\nsd=2Jsd\n/planckover2pi1ǫνβαsνSβ. (12)\nWenotethat thedefinition ofspincurrent depends onthat ofthes pin relaxationtorque37–39.\nFor example, we consider the case when the spin relaxation torque c an be described by\nτα=Tα+∇iPα\ni, where the polarization Pα\niis an arbitrary vector with ∇iPα\ni= 0, whose\nindexiandαrepresent the direction of the polarization in the real space and th at of the\nspin in the spin space, respectively. Then, the spin current Jα\nican be also represented by\nJα\ni=jα\ni+Pα\niandJα\nisatisfies the conservation law as ∂tsα+∇iJα\ni=τα. We discuss the\nspin current defined in Eq. (9). To consider the spin current and th e spin relaxation torque,\nwe calculate the charge density and spin density in the following subse ctions.\nB. Charge density\nFirst, we will calculate the charge density ρein the linear response to S.ρeis de-\nscribed by using the lesser component of the Keldysh-Green’s func tion,−i/planckover2pi1G<(x,t,x,t) =\n∝an}b∇acketle{tψ†(x,t)ψ(x,t)∝an}b∇acket∇i}htin the same position and time as\nρe=−i/planckover2pi1etr/bracketleftbigˆG<(x,t,x,t)/bracketrightbig\n. (13)\n7Hence,ρeis given by\nρe=i/planckover2pi1eJsd\nL2/summationdisplay\nq,Ωei(Ωt−q·x)tr[ˆΠ0ν(q,Ω)Sν\nq,Ω], (14)\nwhereL2is the area of the disordered surface of the TI, and q= (qx,qy) and Ω indicate\nthe momentum and frequency of the localized spin Sν\nq,Ω(ν=x,y,z), respectively. Here, the\ncharge-spin correlation function ˆΠ0νis given by\nˆΠ0ν(q,Ω) =/summationdisplay\nk,ω[ˆgk−q\n2,ω−Ω\n2ˆΛν(q,Ω)ˆgk+q\n2,ω+Ω\n2]<, (15)\nwhere ˆgk±q\n2,ω±Ω\n2is the non-perturbative Green’s function of HTIincludingVi, which is taken\ninto account within the Born approximation. The retarded (advanc ed) Green’s function ˆ gr\nk,ω\n(ˆga\nk,ω= [ˆgr\nk,ω]†) is given by\nˆgr\nk,ω= [/planckover2pi1ω+ǫF−/planckover2pi1˜vFˆσ·(k×z)+i/planckover2pi1/(2τ)]−1,\nwhere/planckover2pi1/(2τ) =πniu2\niνe/2 represents the self-energy due to Viwithin the Born approxima-\ntion. The vector ˆΛνin Eq. (15) is the vertex function, which is described by\nˆΛγ(q,Ω) = ˆσγ+/summationdisplay\nζ=0,x,y,z[˜Γ(q,Ω)+[˜Γ(q,Ω)]2+···]γζˆσζ. (16)\nHere ˆσ0=ˆ12×2is the identity matrix and ˜Γγζis given by ˆΓγ:\nˆΓγ(q,Ω)≡niu2\ni/summationdisplay\nkˆgk−q\n2,ω−Ω\n2ˆσγˆgk+q\n2,ω+Ω\n2\n=˜Γγζˆσζ. (17)\nThe correlation function ˆΠ0νcan be decomposed into the retarded and advanced Green’s\nfunction by using the formula ˆ g<\nk,ω=fω(ˆga\nk,ω−ˆgr\nk,ω)40, wherefωis the Fermi distribution\nfunction. Using the formula, we can estimate the correlation funct ionˆΠ0νon the surface of\nthe doped TI, i.e.,/planckover2pi1/(ǫFτ)≪1 regime as ˆΠ0γ=ˆΠra\n0γ+o(/planckover2pi1/(ǫFτ)), where ˆΠra\n0γis represented\nby\nˆΠra\n0γ=/summationdisplay\nk,ω(fω+Ω\n2−fω−Ω\n2)ˆgr\nk−q\n2,ω−Ω\n2ˆΛra\nγˆga\nk+q\n2,ω+Ω\n2. (18)\nHereˆΛra\nγis given by the Pauli matrix as\nˆΛra\nγ=/summationdisplay\nζ=0,x,y,z/bracketleftbigˆ1+˜Γra+(˜Γra)2+···/bracketrightbig\nγζˆσζ≡/summationdisplay\nζ˜Λra\nγζˆσζ, (19)\nˆΓra\nγ=niu2\ni/summationdisplay\nkˆgr\nk−q\n2,ω−Ω\n2ˆσγˆga\nk+q\n2,ω+Ω\n2. (20)\n8Using Eqs. (19)-(20) under the condition Ω τ≪1, we can calculate ˆΠra\n0γin the low-\ntemperature limit. Besides this, we can calculate the response func tionˆΠra\n0γby postuating\nΩτ≪1,qℓ≪1, andq2\nx=q2\ny=q2/2.ˆΠra\n0γis given by\nˆΠra\n0γ=−Ω\n2π/summationdisplay\nζ=0,x,y,zˆΓra\nζ˜Λra\nγζ, (21)\nwhereˆΓra\nζ≡/summationtext\nν=0,x,y,z˜Γra\nζνσνcan be expressed by 4 ×4 matrix ˜Γraas\n˜Γra=\n1−iΩτ−1\n2ℓ2q2 i\n2ℓqy −i\n2ℓqx 0\ni\n2ℓqy1\n2(1−iΩτ−1\n2ℓ2q2)1\n4ℓ2qxqy−i\n4qxℓ/planckover2pi1\nǫFτ\n−i\n2ℓqx1\n4ℓ2qxqy1\n2(1−iΩτ−1\n2ℓ2q2)−i\n4qyℓ/planckover2pi1\nǫFτ\n0i\n4qxℓ/planckover2pi1\nǫFτi\n4qyℓ/planckover2pi1\nǫFτo(/planckover2pi1\nǫFτ)2\n.(22)\nIn the above equation, we have used niu2\niπνe/(/planckover2pi1/2τ) = 1/2. Here,νe=ǫF/(2π/planckover2pi12˜v2\nF) is\nthe density of states at the Fermi energy on the surface of the d oped TI. From the above\nequation, themagnitudesof ˜Γζzand˜Γzζarenegligiblysmaller thanthatof ˜Γνµ(ν,µ= 0,x,y)\nfor/planckover2pi1/(ǫFτ)≪1. As a result, ˆΓµ=/summationtext\nν=0,x,y[˜Γ]µνˆσν+o(/planckover2pi1/(ǫFτ)) is obtained by\nˆΓra\n0=/parenleftbigg\n1−iΩτ−1\n2ℓ2q2/parenrightbigg\nˆσ0+i\n2ℓˆσaqbǫabz, (23)\nˆΓra\nµ=x,y=/bracketleftbigg1\n2/parenleftbigg\n1−iΩτ−3\n4ℓ2q2/parenrightbigg\nδµν+1\n4ℓ2qµqν/bracketrightbigg\nˆσν+i\n2ℓqaǫµazˆσ0. (24)\nThen,˜Λra\nζγcan also be estimated by using ˜Γraas41\n˜Λra\nγζ= [(1−˜Γra)−1]γζ. (25)\nTherefore, from Eqs. (13) and (21)-(24), the charge density ρeis obtained by\nρe=eνeJsdτ\nL2/summationdisplay\nq,Ωei(Ωt−q·x)ℓΩ\nq2ℓ2+iΩτ(qySx\nq,Ω−qxSy\nq,Ω)\n=−eνeJsdτℓ[∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z, (26)\nFrom Eq. (26), we find that the charge density ρeis induced by ∂t[∇× ∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z. Here\n∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htDis defined by the convolution of the in-plane of the localized spin S/bardbland a diffusion\npropagator function Don the disordered surface of the TI as\n∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD(x,t)≡1\nτ/integraldisplay\ndt′/integraldisplay\ndx′D(x−x′,t−t′)S/bardbl(x′,t′), (27)\nD(x,t) =1\nL2/summationdisplay\nq,Ωei(Ωt−q·x)1\n2Dq2+iΩ, (28)\n9where,D≡˜v2\nFτ/2 is a diffusion constant and ∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htDdenotes the nonlocal spin, which\ndiffusively propagates by the diffusion propagator D. TheDresults because of nonmagnetic\nimpurity scattering on the disordered surface of the doped TI. We also find that the charge\ndensity due to the out-of plane of the localized spin, Sz, is negligible smaller than that due\ntoS/bardbl.\nThe diffusion propagator Dsatisfies the differential equation\n(∂t−2D∇2)D(x−x′,t−t′) =δ(x−x′)δ(t−t′). (29)\nWe find that from Eqs. (26) and (27), the diffusive motion of the cha rge density obeys the\ndiffusion equation:\n(∂t−2D∇2)ρe=−eνeJsdℓ(∇×∂tS/bardbl)z. (30)\nThe above equation means that the diffusion propagator of the cha rge density is caused\nby the spatial and time derivative of the localized spin, ( ∇×∂tS/bardbl)z, on the surface of the\ndoped TI. When the localized spin is spatially uniform, ρeis not driven by the magnetization\ndynamics.\nC. Spin current\nWe will now consider the spin current due to the magnetization dynam ics on the disor-\ndered surface of the doped TI. The spin current is proportional t o the charge density [see\nEq. (9)]. From the result of the charge density due to magnetizatio n dynamics [see Eq.\n(26)], the spin current is given by\njα\ni=−1\n2ǫzαiνeJsdℓ2[∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z. (31)\nThis is one of the main results of this paper. From Eq. (31), the direc tion of spin of the\nspin current is perfectly locked and is perpendicular to the direction of the flow of the spin\ncurrent. The origin lies on the spin-momentum locking on the surface of the TI. The spin\ncurrent is proportional to the coefficients, which are the density o f states at Fermi energy\nνe, thes-dexchange coupling Jsd, and the square of the mean-free path ℓ2. Herejα\niis\nproportional to the spatial and time derivative of the nonlocal spin as [∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z. We\nfind that the local spin does not contribute to the spin current gen eration. In the case\n10when the spin structure of the MI is spatially uniform, the spin curre nt vanishes. Since the\nspin current is proportional to the charge density, we expect tha t the spin current can be\narising from the accumulation of the diffusive charge density, which is given by Eq. (30).\nAdditionally, Eq. (31) indicates that the spin current is an even-fun ction of ˜vF, the sign of\nwhich depends on the helicity of electron on the surface of the TI. T herefore, the direction\nof the spin and flow of the spin current on top surface ( jα\ni,top) and that on bottom surface of\nthe TI (jα\ni,bottom) are equal as jα\ni,top=jα\ni,bottom. We find that the in-plane component of the\nlocalized spin S/bardbl≡S−Szzcontributes to the spin current, but the out-of plane component\nof the localized spin Szzdoes not. We expect that its origin lies on the spin-orbit coupling\nofHTI. From ( ˆσ×p)z=ˆσ·(p×z) andS/bardbl= (z×S/bardbl)×z, the Hamiltonian HTI+Hsd\ncan be described by\nHTI+Hsd=/integraldisplay\ndxψ†/braceleftbigg\n˜vFˆσ·/bracketleftbigg/parenleftbigg\np−Jsd\n˜vF(z×S/bardbl)/parenrightbigg\n×z/bracketrightbigg\n−JsdSzˆσz−ǫF/bracerightbigg\nψ(32)\nFrom the above equation, we can regard that the conduction elect rons momentum pis\nshifted by the in-plane localized spin S/bardbl:p→p−Jsd\n˜vF(z×S/bardbl). The in-plane localized spin\nz×S/bardblplays a role like an electromagnetic vector potential A=Jsd\ne˜vF(z×S/bardbl)13,14. Then, the\nobservable quantity should be proportional to the gauge invariant form: an effective electric\nfieldE≡ −∂tAor an effective magnetic field B=∇×A, as represented by\nE=−Jsd\ne˜vF(z×∂tS/bardbl), (33)\nB=Jsd\ne˜vF∇×(z×S/bardbl). (34)\nThedynamics ofthein-planecomponent ofthelocalizedspincanbere gardedastheeffective\nelectromagnetic field, which acts as a driving force to trigger the mo tion of conduction\nelectrons. While, theout-ofplaneone Szzplays arolelike magnetic fields fortheconduction\nelectronsanddoesnotdirectlyshift pinthemomentumspace. Weexpectfromthedifference\nof these properties of the localized spin, the contribution from Szzcould be smaller than\nthat from S/bardbl.\nThe spin current can be represented by using the effective electric fieldEand{∇×\n[∝an}b∇acketle{tE∝an}b∇acket∇i}htD×z]}z=−∇·∝an}b∇acketle{tE∝an}b∇acket∇i}htDas\njα\ni=−1\n2ǫzαiνee˜vFℓ2∇·∝an}b∇acketle{tE∝an}b∇acket∇i}htD. (35)\n11From the above equation, we find that the spin current is proportio nal to∇·∝an}b∇acketle{tE∝an}b∇acket∇i}htDstemming\nfrom charge density. In fact, the charge density can be represe nted byρe∝∇· ∝an}b∇acketle{tE∝an}b∇acket∇i}htD, as\nshown in Eq. (26). Here, the charge density is also proportional to ∇·∝an}b∇acketle{tE∝an}b∇acket∇i}htDand is similar to\nthe Gauss’s law in Maxwell equations as ρe=ǫ∇·E, whereǫis a permittivity and Eis an\napplied electric field. Thus, we can interpret that the charge densit y and the spin current on\nthe surface of the TI are generated by the divergence of the effe ctive electric field. Equations\n(26), (31), and (35) are the main results of this section.\nIV. CHARGE CURRENT DUE TO MAGNETIZATION DYNAMICS\nIn this section, we show charge current due to magnetization dyna mics on the disordered\nsurface of the doped TI. Because of the spin-momentum locking, t he charge current is\nproportional to the density of the spin polarization on the surface of the doped TI. We\ncalculate the spin density, the charge current and the resulting sp in-relaxation torque.\nA. Spin density\nTo discuss the charge current, we calculate the spin density due to the magnetization\ndynamics in the linear response to the localized spin. The spin density s=1\n2∝an}b∇acketle{tψ†ˆσψ∝an}b∇acket∇i}htis\ngiven by\nsµ=i/planckover2pi1Jsd\n2L2/summationdisplay\nq,Ωei(Ωt−q·x)tr[ˆΠµν(q,Ω)Sν\nq,Ω], (36)\nwhere, Π µν(µ,ν=x,y,z) is the spin-spin correlation function. Π µνcan be calculated within\nthesameformalismasinthesection3.2,andisrepresentedby ˆΠµν= ˆσµˆΠ0ν. Fromtheresult,\nwe can obtain the spin density s. Here,scan be decomposed into two terms: s=s/bardbl+szz,\nwheres/bardbl=s−szzandszzshow the in-plane and out-of plane component of the spin\non the disordered surface of the doped TI, respectively. We find t hatszzis proportional\nto∂tSz, and its magnitude is negligibly smaller than that of the magnitude of s/bardblwithin\nthe approximation |sz|/|s/bardbl| ∼o[/planckover2pi1/(ǫFτ)≪1]. Thus, the spin density can be estimated\nbys=s/bardbl+o[/planckover2pi1/(ǫFτ)] andSzdoes not contribute to the generation of s. The dominant\ncontribution of scan also be decomposed into two terms:\ns=sL+sD, (37)\n12wheresLis the local spin density due to S/bardbl, and is given by\nsL=−1\n2νeJsdτ∂tS/bardbl. (38)\nThe local spin density sLis induced by the time-derivative of the in-plane component of the\nlocalized spin ∂tS/bardbl. On the other hand, the second term of Eq. (37), sD, is the diffusive\nspin density and is given by\nsD=−1\n2νeJsdτℓ2(z×∇)[∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z (39)\n=ℓ\n2e(z×∇)ρe. (40)\nFrom Eq. (40), sDis generated by the driving field ( z×∇)(∇×∂t∝an}b∇acketle{tS∝an}b∇acket∇i}htD)z, which is the\nspatial gradient and the time-derivative of the nonlocal localized sp in∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD. The driving\nfield is also described by ( z×∇)(∇×∂t∝an}b∇acketle{tS∝an}b∇acket∇i}htD)z=∂t[∇2∝an}b∇acketle{tS∝an}b∇acket∇i}htD−∇(∇· ∝an}b∇acketle{tS∝an}b∇acket∇i}htD)]. Here, sD\nis described by the spatial gradient of the charge density, which is c aused by the electron\ndiffusion on the surface of the TI, as shown in Eq. (26). In addition, sDis also represented\nby the spin current: The charge density is proportional to the spin current,ρe=e\n˜vF(jx\ny−jy\nx),\nandsDbecomes\nsD=1\n2τǫzαi(z×∇)jα\ni. (41)\nFrom Eqs. (38)-(39), we find that sis an even-function of ˜ vFand is independent of the\nhelicity on the surface of the TI. Therefore, the direction of sdoes not depend on whether\nwe are focusing on the top or bottom surface of the TI.\nWe find that from Eqs. (38)-(39), the spin is polarized not by a stat ic magnetization\nbut by magnetization dynamics. Therefore, we expect that static magnetization does not\ninduce spin polarization on the surface of the TI. This seems to be an omalous property on\nthe surface. The response between the spin polarization and the s tatic magnetization on\nthe surface of the TI is different from that in conventional metals: In the metals, a spin is\npolarized even by static magnetization. We will consider shortly why s tatic magnetization\ndoes not generate spin polarization on the surface of the TI. The m agnetization on the\nsurface of the TI plays the role to shift the momentum of conductio n electrons from pinto\np−Jsd\n˜vF(z×S/bardbl) in momentum space. As a result, the center of Fermi sphere is also shifted\nfromp= 0 into p=Jsd\n˜vF(z×S/bardbl). Then, the direction of the spin at each momentum\nare perfectly perpendicular to that of the momentum. Besides, th e spin configuration in\n13momentum space does not change before and after the shift, bec ause the direction of the\nspin at each momentum are independent of the shift. Therefore, n o spin polarization is\ndriven by momentum shift due to a static magnetization.\nB. Charge current\nWe note that on the surface of the TI, the charge current jis proportional to the spin\ndensity on the surface of the TI. The charge current is represen ted by using the renormarized\nvelocity operator as j= 2e˜vF(z×s)36. From Eqs. (38)-(40), the charge current is also\ndecomposed into two terms: j=jL+jDas\njL=−eνeJsdℓ(z×∂tS/bardbl), (42)\njD=eνeJsdℓ3∇[∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z. (43)\nThejLis the local charge current and is induced by the time-derivative of t he localized spin\nS/bardbl. The direction of jLis parallel to z×S/bardbl. On the other hand, jDis the diffusive charge\ncurrent caused by impurity scattering on the disordered surface of the TI. In fact, jDcan be\nrepresented by the spatial gradient of the charge density as jD=−2D∇ρe. This means that\ndiffusive current is generated by the spatial gradient of the charg e density on the surface of\nthe TI. We note that the charge current is an odd-function of ˜ vF, so that the direction of\nthe charge current on the top surface jtopis opposite to that on the bottom surface jbottom,\nifSis same on the top and bottom surface. It is noted that, from Eqs. (26) and (42)-(43),\nthe charge density ρeand charge current jsatisfy the conservation law ˙ ρe+∇·jγ= 0. The\ndetail is shown in Appendix B.\nNext, we comment on the relationship between the spin current and the charge current\non the disordered surface of the doped TI. Substituting ǫzαijα\ni= ˜vFρe/einto Eq. (43), we\nfind that the diffusive charge current can be described by the spin c urrent as\njD=−eℓǫzαi∇jα\ni. (44)\nThis is also the main result of this paper. We expect that the above eq uation displays\nthe conversion between the spin current into the diffusive charge c urrent on the disordered\nsurface of the doped TI by using the spatial gradient of the spin cu rrent. The spin current\ncan be converted into the diffusive charge current when the spin cu rrent depends on the\n14space on the disordered surface. The relation in Eq. (44) is plausible on the disordered\nsurface of the doped TI, because the charge density ρeis proportional to the spin current,\nand a diffusive particle current generally proportional to a spatial g radient of particles. We\nnote that there is no relation between the spin current and the loca l charge current jL.\nC. Effective conductivity\nThechargecurrent duetomagnetizationdynamics jcanbealsodescribed bytheeffective\nelectric field E:\nj=e2˜v2\nFνeτE+e2˜vFνeℓ3∇[∇·∝an}b∇acketle{tE∝an}b∇acket∇i}htD]. (45)\nThe first term and the second term are corresponding to the local and diffusive charge cur-\nrent, respectively. From the above results, we will consider an effe ctive conductivity: an\nefficiency of the charge flow due to the applied effective electric field E. This is similar\nto the conventional electric conductivity: In general, the longitud inal electrical conductiv-\nity is defined from dividing the charge current by an applied electric fie ld42. We expect\nthat the current corresponds to the local current. Then, an eff ective longitudinal electrical\nconductivity can be defined by jL=σE, and is given by the first term in Eq. (45) as\nσ=e2˜v2\nFνeτ. (46)\nTheconductivityonlydependsontheparametersonthesurfaceo ftheTI,andisindependent\nof parameters attached to the MI.\nD. Spin relaxation torque\nWe will consider the spin relaxation due to the magnetization dynamics on the surface\nof the TI. Using Eqs. (31) and (37)-(39), we can describe ∂tsαand∇ijα\nias\n∂tsα=−1\n2νeJsdτ∂2\ntS/bardbl,α+1\n2νeJsdℓ2τǫαiz∇i[∇×∂2\nt∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z, (47)\n∇ijα\ni=−1\n2νeJsdℓ2ǫαiz∇i[∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z. (48)\nTherefore, the spin relaxation torque Tα=∂tsα+∇ijα\niin the linear response to S/bardblis\nobtained by\nT=−1\n2νeJsdτ∂2\ntS/bardbl+1\n2νeJsdℓ2(1−τ∂t)(z×∇)[∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z+O(J2\nsd).(49)\n15The first term in Eq. (47) shows a local spin relaxation torque and is in duced by∂2\ntS/bardbl. The\nsecond term is a nonlocal spin relaxation torque and is driven by ( z×∇)(∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD)z.\nThe nonlocal torque vanishes when the magnetization is spatially unif orm. The third term\nin Eq. (47) represents the higher order of JsdandS. From Eq. (12), the spin relaxation\ntorqueTsdis proportional to Sands, which is proportional to S/bardbl. Therefore, the third\nterm in Eq. (47) corresponds to Tsdwithin the linear response to S. We expect that the\nthird term Tsd≃o(Jsdτ\n/planckover2pi1)2can be negligibly small and be ignored in comparison with the\nfirst and the second terms of Tin the regimeJsdτ\n/planckover2pi1≪1.\nThe spin relaxation torque Tis also represented by the effective electric field as\nT=1\n2eνe˜vFτ(∂tE×z)+1\n2eνe˜vFℓ2(1−τ∂t)(z×∇)[∇·∝an}b∇acketle{tE∝an}b∇acket∇i}htD]. (50)\nThe local spin relaxation is written as the time-derivative of the effec tive electric field and\nthe diffusive one is induced by the spatial gradient of the nonlocal eff ective electric field. The\nelectric field dependence of the spin relaxation torque on the surfa ce of the TI is different\nfrom that in NM with spin-orbit interactions: The spin relaxation torq ue in the NM, TNM,\nis proportional to the spatial gradient of the applied electric field39. We expect that the\ndifferencecanbecausedbythe k-dependenceoftheenergydispersion: Theenergydispersion\non the surface of the TI is a linear function of k, while that in the NM proportional to the\nsquare of k. Equations (39), (44), and (46) are the main results of this sectio n.\nV. DISCUSSION\nA. Spin torque\nWe will phenomenologically study the change of the magnetization dyn amics in before\nand after the spin-charge generation due to the ferromagnetic r esonance (FMR). Now, we\nconsider a disordered surface of the MI/TI junction, as shown in F ig.1. Further, in the\njunction, a static magnetic field and ac magnetic field are additionally a pplied. Here ac\nmagnetic field is given by a microwave irradiation, and is needed for FMR in the MI. When\nwe apply this magnetic field, the magnetization dynamics is triggered in the MI, and the\nmagnetization dynamics induces the spin polarization on the surface of the TI [see Eqs.\n(37)-(39)]. Then, the induced spin polarization splays the role of an exchange field acting\non the magnetization in the MI. As a result, the magnetization dynam ics is affected from\n16the generated spin s. Here, the mutual interaction between the magnetization dynamic s\nand the induced spin are called as the feedback effect25.\nThe magnetization dynamics on the disordered surface of the dope d TI is described from\nthe Landau–Lifshitz–Gilbert (LLG) equation as43\n∂tM=−γµ(M×H)+α\nMM×∂tM+Te, (51)\nwhere,M=−M\nSSis the magnetization of the MI, γis the gyromagnetic ratio, µis a\npermeability, αis a Gilbert damping constant, H=H0+Hacis an applied magnetic field\non the disordered surface of the doped TI. H0andHacdenote a static and ac magnetic\nfield, respectively. The spin torque on the disordered surface of t he doped TI is given by\nTe=2Jsda2\n/planckover2pi1M×s. The torque can be decomposed into two terms: Te=TL\ne+TD\ne,\nwhereTL\ne=2Jsda2\n/planckover2pi1M×sLandTD\ne=2Jsda2\n/planckover2pi1M×sDare the local and diffusive spin torque,\nrespectively. Here, ais a lattice constant on the surface of the TI. These spin torques a re\nobtained from Eqs. (37)-(39) and S/bardbl=−S/bardbl\nM/bardblM/bardblas\nTL\ne=κ\nM/bardblM×∂tM/bardbl, (52)\nTD\ne=κ\nM/bardblℓ2M×(z×∇)(∇×∂t∝an}b∇acketle{tM/bardbl∝an}b∇acket∇i}htD)z, (53)\nwhere,M/bardblis the magnitude of the in-plane magnetization M/bardbl≡M−Mzz, andκ=\nνea2J2\nsdτS/bardbl//planckover2pi1is dimensionless coefficient proportional to J2\nsdandτ. We find that TL\ne∝\nM×∂tM/bardblis slightly different from the damping torqueα\nMM×∂tM; which is a damping\nof the magnetization. We could expect that the contribution from t he local spin torque TL\ne\ncan be observed in the experiments on the surface of FM/TI junct ion27,28. Here,TL\neplays\nthe role of an anisotropic damping torque unless the static magnetic field and microwave\nare parallel to the zdirection. The anisotropic damping affects the magnetic permeability :\nFor example, when the static magnetic field and microwave are paralle l to they-direction,\nthen, the longitudinal magnetic permeability χxxandχzzare not equal each other. Here,\nTD\neseems to be a new type of spin torque TD\neon the surface of the TI. TD\nein Eq. (53) is\ninduced by the spatial gradient of the magnetization, M×(z×∇)(∇×∂t∝an}b∇acketle{tM/bardbl∝an}b∇acket∇i}htD)z. When\nthe magnetization is spatially uniform, TD\neis zero and TL\neis nonzero.\nSince,jα\niandTD\neare proportional to the charge density on the disordered surfac e of the\ndoped TI, TD\necan be described by the spin current jα\ni:\nTD\ne=Jsdτa2\n/planckover2pi1M×[(z×∇)ǫzαijα\ni]. (54)\n17The spatial gradient of the spin current induces the diffusive spin to rque on the disordered\nsurface of the doped TI. We expect that the contribution of jα\niis detected from the change\nof the half-width value, as well as the change of a shift of the magne tic resonance frequency\nthroughTD\ne[discussed in Sec.VC]. Equations (52)-(54) are the main results of th is section.\nB. Magnetic permeability without diffusion\nUsing Eqs. (51)-(53), we discuss the magnetic permeability in FMR wh en the magneti-\nzation is spatially uniform on the surface of the MI/TI junction. We c onsider that in the\njunction, the applied static magnetic field H0and the microwave of the ac magnetic field\nHacare applied along the ydirection: H0= (0,H,0) andHac= (hx,0,hz). For an uniform\nmagnetization case, the spin becomes sL∝ne}ationslash= 0 and sD= 0, and the spin torque are TL\ne∝ne}ationslash= 0,\nbutTD\ne= 0.\nThen, the LLG equation on the surface of the MI/TI junction can b e described by\n∂tM=γµ(H×M)+α\nM(M×∂tM)+κ\nM/bardbl(M×∂tM/bardbl). (55)\nTo estimate the magnetic permeability on the surface of the TI, we a ssume that the |H0|is\nlarger than the |Hac|,i.e.,|hx|≪Hand|hz|≪H. Then, from the applied magnetic field,\nwe expect that the local magnetization on the surface M= (mx,My,mz) can be satisfied\nmx≪Myandmz≪My. Moreover we assume that the time-dependence of the precessio n\nofMis given by mx∝mz∝eiΩtand∂tMy∼0. In order to solve the LLG equation, we\ntake a linear approximation of mi:mimj∼0,My∼M(=|M|), andM/bardbl∼M. Then, the\nLLG equation becomes\n∂tmx=γµ(Hmz−hzM)+α∂tmz,\n∂tmz=γµ(hxM−Hmx)−(α+κ)∂tmx,(56)\nand the magnetic permeability is given by\n/parenleftbiggmx\nmz/parenrightbigg\n=/parenleftbiggχxxχxz\nχzxχzz/parenrightbigg/parenleftbigghx\nhz/parenrightbigg\n. (57)\nThe frequency dependence of the longitudinal magnetic permeabilit yχxxandχzzare de-\n18scribed by\nχxx=ωM(ωH+iαΩ)\n(ωH+iαΩ)[ωH+i(α+κ)Ω]−Ω2, (58)\nχzz=ωH+i(α+κ)Ω\nωH+iαΩχxx, (59)\nwhere.ωH=γµHandωM=γµMare the angular frequency of the applied static magnetic\nfield and that of the magnetization, respectively. The permeability χxxandχzzare different\neach other originating from the anisotropic spin-transfer torque TL\ne. Transverse magnetic\npermeability has the relation χxz=−χzxand is given by\nχxz=−iΩ\nωH+iαΩχxx. (60)\nThe real part of the permeability Re[ χxx] shows a Lorentzian profile due to the magnetic\nresonance around the resonant frequency Ω r, which is proportional to H. The Im[χxx]\nindicates the energy absorption of the applied microwave around Ω r. Half-width value of\nIm[χxx] expresses the damping of the precessional motion of the magnet ization. From Eq.\n(58), the half-width value ∆Ω is caused by the Gilbert damping ( αM×∂tM) and the local\nspin torque TL\negiven by\n∆Ω≃(2α+κ)Ωr. (61)\nWe expect that in the MI without TI, the half-width value of the magn etic permeability\nestimates ∆Ω = 2 αΩr. Eq. (61) indicates that on the surface of the doped TI, ∆Ω is\nenhanced from 2 αΩrinto (2α+κ)Ωr. The origin lies on TL\ne, which is triggered by the\ninduced spin, where the spin is induced by the magnetization dynamics on the surface of\nthe doped TI. This enhancement of ∆Ω has been verified in the recen t experiment28.\nWe compare ∆Ω in MI/TI junction with ∆Ω in FM/NM junction. In the FM/ NM\njunction, it has been demonstrated that the enhancement of ∆Ω is triggered by the spin\ncurrent in the NM, which is generated by the magnetization dynamics of the FM44. On the\nsurface of the MI/TI junction with uniform spin structure, on the other hand, spin current\nis not generated by magnetization dynamics, and the spin current d oes not contribute to\n∆Ω. The enhancement of ∆Ω is caused by the spin polarization due to t he magnetization\ndynamics on the surface.\n19C. Magnetic permeability with diffusion\nWe will consider the surface of the TI/MI junction, where the localiz ed spin in the MI is\nspatially inhomogeneous. The spin structure of the localized spin we c onsider is a spin-wave\nor longitudinal conical spin structure, which are realized in ferrimag netic insulator yttrium\niron garnets (YIG) or multiferroics45, respectively. We expect that even on the surface of\nthe TI, the localized spin depends on the position through proximity e ffects from the MI.\nThen, if we apply ac magnetic field of microwave in the junction along th eydirection, we\nassume that the localized spin becomes precessional motion by the a pplied magnetic field.\nThe localized spin on the surface of the TI can be described by\nS(x,t) = [Scos(q·x−Ωt),Sy,Ssin(q·x−Ωt)], (62)\nwhere,SandSyare a constant coefficient independent of space. We assume |S|≪|Sy|\nand|S|∼Sy. Hereq= (qx,qy) is the momentum of the localized spin and is assumed to\nbe monochromatic. The direction and the magnitude of qdepends on materials of the MI.\nIn the spin structure, nonlocal diffusive spin ∝an}b∇acketle{tS∝an}b∇acket∇i}htDis given by Eqs. (27) and (62) as\n∝an}b∇acketle{tSx∝an}b∇acket∇i}htD=Aq,ΩScos[q·x−Ωt]−Bq,ΩSsin[q·x−Ωt],\n∝an}b∇acketle{tSz∝an}b∇acket∇i}htD=Bq,ΩScos[q·x−Ωt]+Aq,ΩSsin[q·x−Ωt].(63)\nThe component of ∝an}b∇acketle{tS∝an}b∇acket∇i}htDis different from that of S:∝an}b∇acketle{tSx∝an}b∇acket∇i}htDhas cos[q·x−Ωt] components,\nas well as sin[ q·x−Ωt] components. Here, the coefficients AandBare obtained (see\nAppendix A) as\nAq,Ω=q2ℓ2\n(Ωτ)2+(q2ℓ2)2, (64)\nBq,Ω=Ωτ\n(Ωτ)2+(q2ℓ2)2. (65)\nThecoefficients Aq,ΩandBq,ΩdependonqℓandΩτ, whereparameters ℓandτaredetermined\nby the TI, and qand Ω are chosen as characteristic values of the MI. Then, the diffu sive\nspinsDis given from Eqs. (39) as\nsD=νeJsdτℓ2[q2∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD−q(q·∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD)]. (66)\nThen, nonlocal diffusive spin torque TD\neis obtained by using ∝an}b∇acketle{tM/bardbl∝an}b∇acket∇i}htD=−(M/bardbl/S/bardbl)∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htDas\nTD\ne=−κℓ2\nM/bardbl/bracketleftbig\nq2M×∂t∝an}b∇acketle{tM/bardbl∝an}b∇acket∇i}htD−(M×q)(q·∂t∝an}b∇acketle{tM/bardbl∝an}b∇acket∇i}htD)/bracketrightbig\n. (67)\n20The first term of Eq. (67) plays the role of an anisotropic damping to rque, which is similar\ntoTL\ne. The direction of this torque is independent of the direction of q. On the other hand,\nthe second term of Eq. (67) is proportional to ( M×q)(q·∂t∝an}b∇acketle{tM/bardbl∝an}b∇acket∇i}htD), and its direction\ndepends on q.\nUsing Eqs. (62)-(67), we consider the magnetic permeability affect ed byTD\ne.Then, the\nLLG equation is given within the linear order of the magnetization as\n∂tmx=ωHmz−hzωM+α∂tmz\n∂tmz=hxωM−ωHmx−(α+κ)∂tmx+κq2\nyℓ2∂t∝an}b∇acketle{tM/bardbl,x∝an}b∇acket∇i}htD.\nThelasttermoftheaboveequationiscausedby TD\ne. Inordertodiscuss thepermeabilitydue\ntoTD\ne, we consider when the momentum has the y-component ( q=qy), whose direction\nis parallel to the applied static magnetic field. That means that the sp in structure we\nconsider is a longitudinal conical spin structure. From the above eq uation, using ∝an}b∇acketle{tM/bardbl,x∝an}b∇acket∇i}htD=\nAq,Ωmx−Bq,Ωmz, we obtain the magnetic permeability as\n/parenleftbiggmx\nmz/parenrightbigg\n=/parenleftbiggχD\nxxχD\nxz\nχD\nzxχD\nzz/parenrightbigg/parenleftbigghx\nhz/parenrightbigg\n. (68)\nHere the longitudinal and transverse magnetic permeability are give n by\nχD\nxx(q,Ω) =(ωH+iαΩ)ωM\n[ωH+iαΩ][ωH+i(α+ ˜κq,Ω)Ω]−ζq,ΩΩ2, (69)\nχD\nzz(q,Ω) =ωH+i(α+ ˜κq,Ω)Ω\nωH+iαΩχD\nxx, (70)\nχD\nxz(q,Ω) =−ζq,ΩχD\nzx. (71)\nThe obtained permeability is different from that in Eqs. (58)-(60). T he difference is caused\nby coefficients ˜ κq,Ωandζq,Ω:\n˜κq,Ω=κ(1−q2ℓ2Aq,Ω), (72)\nζq,Ω= 1+κq2ℓ2Bq,Ω. (73)\nThe ˜κq,Ωandζq,Ωdepend on qℓand Ωτ. Ifq= 0, one can demonstrate ˜ κq,Ω=κand\nζq,Ω= 1, and the magnetic permeability in Eqs. (69)-(71) equal to that in Eqs. (58)-(60),\nrespectively. Figure 2(a) indicates the Ω τdependence of ˜ κ/κfor several momentum qℓ. The\nparameter ˜κ/κapproaches to 0 from ˜ κ/κ= 1 with increasing Ω τ: In the case for ˜ κ/κ= 0,\n21FIG. 2: (Color online) (a) The Ω τdependence of ˜ κ/κfor several qℓ. (b) The qℓdependence of the\nζfor several frequency (2 πf= 1,2,5,10 GHz) in the fixed relaxation time ( τ= 0.087 ps).\nTL\neandTD\neare canceled out each other, and the spin torque Tevanishes. On the other\nhand, ˜κ/κ= 1 means that TD\neis zero and TL\neis nozero. The relation ˜ κ/κsignificantly\nchanged when ( qℓ)2∼Ωτis satisfied. Figure 2(b) shows the qℓdependence of ζfor several\nangular frequency of the applied ac magnetic field Ω, where we take a realistic relaxation\ntime (τ= 0.087 ps46). The magnitude of ζchanges around qℓ∼√\nΩτand approaches to\nζ= 1 with increasing qℓ.\nFigure 3 (a) and (b) show the Ω τdependence of the real and imaginary part of the\nlongitudinal magnetic permeability, Re[ χD\nxx] and−Im[χD\nxx], respectively. In Figs. 3(a) and\n(b), we choose realistic parameters of the TI: ℓ= 40nm,vF= 4.6×105m/s46,τ= 0.087 ps,\nξ= 0.003,and ˜vF/vF= 0.997. Besides, wechoosematerialparametersoftheferromagne ts28:\nα= 0.015,Jsd∼6meV,S/bardbl∼0.3, and Fermi wavenumber kF= 3.9×108m−1. Then,\nκ= 0.008 is obtained. The frequency of the magnetization ωM/(2π) =fM= 0.28 GHz\nin the MI. The magnitude of the frequency is evaluated by the mater ial parameters of the\npermalloy28.\n22FIG. 3: (Color online) (a)-(b) The frequency dependence of t he real part and imaginary part of the\nlongitudinal permeability, Re[ χD\nxx] and−Im[χD\nxx], for a fixed Gilbert dampingconstant ( α= 0.015),\nthe relaxation time [ τ= 0.087ps], the anisotropic damping constant [ κ= 0.008], and the frequency\n[fH= 1.6 GHz and fM= 0.28 GHz] for several momentum qℓ.\nWe plot the Re[ χD\nxx] and−Im[χD\nxx] functions as the frequency of the ac magnetic field for\nseveral momentum of the localized spin, when we take the frequenc y of the static magnetic\nfieldωH/(2π) =fH= 1.6 GHz. The magnitude of Ω τcan be estimated as Ω τ∼ωHτ=\n8.6×10−3. Forqℓ <0.03, the permeability Re[ χD\nxx] and−Im[χD\nxx] are not dramatically\nchanged from that without diffusion, Re[ χxx] and−Im[χxx], respectively. The reason is due\nto the profile of ˜ κ/κandζ: Both ˜κ/κandζare about 1 within qℓ<0.03 in Ωτ∼8.6×10−3.\nWhileqℓis nearqℓ= 0.03,χD\nxxchanges: The magnitude of χD\nxxincreases from that of χxx.\nBesides, the resonant frequency of χD\nxxincreases from that of χxx[seeqℓ= 0.03 in Figs. 3\n(a)-(b)]. The change of the resonant frequency is also shown in Fig . 4 (a). We find that\nwhen (qℓ)2∼Ωτis satisfied, ˜ κ/κandζdeviate from 1, as shown in Figs. 2(a)-(b). After\nincreasingqℓfromqℓ= 0.03, the shifted resonant frequency gets back again. The magnitu de\nof the permeability increases with increasing qℓ.\nThe half-width value is also changed by the magnitude of Im[ χD\nxx] for several momentum\n23FIG. 4: (Coloronline) The qℓdependenceoftheresonant frequency(a) andthehalf-width value(b)\nnormalized by the angular frequency of the applied magnetic field for several frequencies (2 πfH=1,\n2, 5, and 10 GHz) in 2 α+κ= 0.038 and 2 α= 0.030.\nqℓ. This trade off between the half-width value and qℓfor several frequency (2 πfH= 1,\n2, 5, and 10 GHz ) is shown in Fig. 4(b). In the case for τ= 0.087 ps, the half-width\nvalue ∆Ω significantly changed around the qℓ, which satisfies qℓ∼√\nΩτ(e.g.,qℓ= 0.03 in\nωH= 10 GHz). Figure 4(a) indicates qℓdependence of the resonant frequency rate Ω r/ωH.\nThe Ω r/ωHchanged from Ω r/ωH∼1 into Ω r/ωH= 0.998 around the qℓ, which satisfies\nqℓ∼√\nΩτ.\nWe discuss the role of diffusive spin torque TD\nefrom Figs. 4(a) and (b). Figure 4(a)\nshows that the resonant angular frequency Ω rtends to decrease with increasing qℓ. The\ndecrease of Ω rcan be caused by the increase of ζq,Ω, and the change of ζq,Ωis caused by TD\ne.\nTherefore, we expect that TD\neplays a role as a field-like torque to shift the resonant fre-\nquency. Then, Fig.4 (b) indicates that the half-width value tends to decrease with increasing\n˜κq,Ωin Eq. (73), which is caused by TD\ne. It means that the damping of the magnetization\ndynamics is reduced by TD\neon the disordered surface of the doped TI. Thus, TD\nebehaves\n24TABLE I: A brief summary of the role of the spin torques and the half-width value in several\nspin structures on the surface of the TI, when the direction o fH0and the propagation of Hacare\nparallel to y-axis.\nUniform Transverse conical ( qy= 0) Longitudinal conical ( qy∝ne}ationslash= 0)\nTL\neDamping torque Damping torque Damping torque\nTD\ne – – Damping torque and field-like torque\n∆Ω/Ωr2α+κ 2α+κ 2α+ ˜κ\nboth the damping torque and the field-like torque. The role of the sp in torque in several\nspin structures are summarized in the Table I. Fromthe table I, we c an expect to distinguish\nthe half-width value contributed from TL\neandTD\neby tuning of the magnitude of the applied\nmagnetic field H0. The reason is why the longitudinal spin structure changes an spat ial\nuniform ferromagnetic structure if when we apply a strong magnet ic field, which broke the\nlongitudinal spin structure. Then, qℓof the spatial uniform spin structure can be regarded\nasqℓ= 0. Therefore, we expect that if when we apply the strong applied m agnetic field\nin the longitudinal spin structure, the half-width value with a finite qℓchanges into the\nhalf-width value with qℓ= 0 as ∆Ω/Ωr= 2α+ ˜κ→2α+κ[see Fig.4 (b)].\nFor magnetization dynamics due to the magnetic resonance, we nee d to apply magnetic\nfields in the MI/TI junction. Then, the spin-charge generation and transport are triggered\nnot only by the magnetization dynamics, but also by the applied magne tic field. We will\nestimate when the contribution due to the magnetic field can be relev ant. The contribution\nfromtheappliedmagneticfieldcanbedescribedbytheZeemaneffect ,HZ=−2/planckover2pi1γ/integraltext\ndxB·s,\nwhereBis the applied magnetic field and couples with conduction electrons spin on the\nsurface of the TI. The contribution from Bcan be treated within the same formalism in\nsections 3 and 4 by replacing S→S+2(/planckover2pi1γ/Jsd)Bin Eq. (3). As a result, the spin-charge\ngeneration and transport due to HZandHsdare obtained by replacing S→S+2(/planckover2pi1γ/Jsd)B\nin Eqs. (26), (31), (38)-(39), and (42)-(43). We expect that t he contribution from HZcan be\nignored compared with that from Hsd, when the energy scale of the Zeeman effect is smaller\nthan that of the exchange energy on the surface of the TI as 2 /planckover2pi1γ|B|/(Jsd|S|)≪1. The\n|B|/Jsdvalue can be estimated by |B|/Jsd≪1/(2/planckover2pi1γ)∼4×104T·eV−1inS∼1. Then, the\n25magnitude of the exchange coupling Jsdcan be estimated. Using realistic parameters, the\nmean-free path ℓ= 40nm, Fermi velocity vF= 4.6×105m/s46, we obtain the /planckover2pi1/τ∼15meV\nandJsd≪15 meV, which is requested for the perturbation condition Jsdτ//planckover2pi1≪1. Then, if\n|B| ≫10T is satisfied, we need to consider the contribution from HZ.\nAt the end of this section, we estimate the spin density due to the dy namics of the\nlongitudinal spin structure S=S/bardbl(sinθcosΩt,1,sinθsinΩt) in the case of θ≪1 around\nresonance angular frequency Ω ∼1×1010s−1. The magnitude of the spin depends on the\nregime ofq2ℓ2≪Ωτorq2ℓ2≫Ωτ. Inq2ℓ2≪Ωτregime, the magnitude of the nonlocal spin\ncan be negligible small compared with that of the local term, and the s pin is estimated by\n|s| ∼1.6×10−10˚A−2atθ∼0.1radintheFMR.Then, wefindthatthemagnitudeofthespin\ndue to the spin-pumping is smaller than that of the spin due to the app lied electric field47.\nOn the other hand, in q2ℓ2≫Ωτregime, the local and nonlocal spin vanishes each other\neven in the presence of FMR. From the results, we expect that the change of the magnitude\nof the spin dependent on q2ℓ2/Ωτcan be measurable for several applied magnetic fields,\nbecause the inhomogenous spin structure ( qℓ∝ne}ationslash= 0) changes into an uniform spin structure\n(qℓ∼0) by using an applied strong magnetic field.\nD. Spin current and charge current\nWewill discuss thespincurrent onthesurfaceofthedisorderedMI /TIjunctioncompared\nwith that in the FM/NM junction. The spin current due to the spin-pu mpingjα\ni,FM/NMin\nthe FM/NM junction is triggered by the magnetization dynamics as1,2,5\njα\ni,FM/NM=b∇i∂tSα+O(S2), (74)\nwherebis a coefficient dependent on materials. It is similar to the spin current in Eq. (74),\nthat thespin current isproportional to a time-dependent magnet izationand thespin current\nvanishes when the magnetization is spatially uniform. The direction of the spin (α) and the\nflow (i) of the spin current in Eq. (74) are not related each other. On the other hand, spin\ncurrent on the surface of the TI, whose direction of spin and flow a re perfectly perpendicular\nto each other. The difference lies on the spin-orbit interaction. The jα\niin Eq. (31) includes\nthe contribution of the spin-orbit interaction, which is absent in Eq. (74) does not.\nCharge current due to the spin-pumping in the FM/NM junction is also given by the\n26magnetization dynamics and Rashba type spin-orbit interactions. W hen the magnetization\nis spatially uniform, the charge current jFM/NMbecomes2,5\njFM/NM=α×(S×∂tS), (75)\nwhereαis a constant vector including the contribution from the spin-orbit in teraction. The\nchargecurrent inEq. (75)isproportionaltothesecond-order o fthelocalizedspin S. That is\ndifferent from the charge current on the surface of TIs. The cha rge current on the surface of\nthe TI is proportional to the localized spin ∂tS/bardbl, as shown in Eq. (43). The difference is due\nto the property of the localized spin: the localized spin plays the role o f the effective vector\npotential on the surface of the TI. We note that the frequency d ependence of these charge\ncurrent is also different; jTI∝∂tS/bardbloscillates with time of the localized spin in the FMR,\nbutjFM/NM∝S×∂tSdoes not. For example, the ac current jTI∝(cosΩt,sinΩt,0) is given\nwhen we apply the magnetic field parallel to the zdirection on the TI/MI junction. On\nthe other hand, the dc current jFM/NM∝Ω(0,0,1) is obtained when we apply the magnetic\nfield parallel to the zdirection in the NM/FM junction.\nVI. SUMMARY\nWe have studied the spin-charge generation and transport due to the magnetization dy-\nnamics on the disordered surface of the doped TI/MI junction. Th e spin current jα\ns,iis\nproportional to the charge density ρeand the direction of its spin and its flow are perfectly\nperpendicular to each other, because of the spin-momentum lockin g on the surface of the\nTI. We have found that jα\niandρeare induced by the time- and spatial-dependent of nonlo-\ncal magnetization dynamics, which is affected by nonmagnetic impurit y scatterings on the\ndisordered surface of the doped TI. These results of jα\niandρeare shown in Eqs. (31) and\n(26), respectively. jα\niandρeis induced except when the magnetization dynamics is spatially\nuniform. We have also shown the induced spin sand charge current density jdue to the\nmagnetization dynamics. Because of the spin-momentum locking, th e spinsand charge\ncurrent density jare proportional to each other. The sandjare generated not only by\nthe local magnetization dynamics, but also by the nonlocal magnetiz ation dynamics with\nthe diffusion on the disordered surface of the doped TI. These res ults ofsandjare shown\nin Eqs. (38)-(39) and (42)-(43), respectively. A brief summary o f the local and nonlocal ρe,\n27TABLE II: A brief summary of the charge, charge current, spin , and spin current density due to\nthe magnetization dynamics on the disordered surface of the TI. These terms are driven by the\neffective electric field E.\ncharge density ρecurrent density jispin density sαspin current density jα\ni\nLocal term – z×∂tS/bardbl∂tS/bardbl–\nNonlocal term [ ∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z∇[∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z(z×∇)[∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z[∇×∂t∝an}b∇acketle{tS/bardbl∝an}b∇acket∇i}htD]z\nDriving field ∇·∝an}b∇acketle{tE∝an}b∇acket∇i}htD E,∇2∝an}b∇acketle{tE∝an}b∇acket∇i}htDz×E, (z×∇)[∇·∝an}b∇acketle{tE∝an}b∇acket∇i}htD] ∇·∝an}b∇acketle{tE∝an}b∇acket∇i}htD\njα\ni,s, andjdue to the magnetization dynamics are represented in Table. II. As a result,\nwe have discussed the modification of the magnetization dynamics be fore and after these\nspin-charge generation and transport on the disordered surfac e of the doped TI. These spin-\ncharge generation and transport can be detected from the half- width value of the magnetic\npermeability in the magnetic resonance in the MI/TI junction, as disc ussed in section V.\nThe magnitude of a Gilbert damping constant αin ferromagnetic insulator is smaller than\nthat in ferromagnetic metals. Then, we can easily detect the chang e of thefHdependence\nof the resonant frequency and half-width value, which are shown in Figs. 4 (a)-(b).\nThe preparation of the hybrid system with the ferromagnetic insula tor deposited on the\nsurface of the TI, EuS/Bi 2Se3, has been reported50, where the magnetic moment of Eu\nlocates at the interface between the EuS and Bi 2Se3. If the magnetization dynamics of\nthe magnetic moment of Eu is triggered by an applied magnetic field, th e spin density\nand charge current can be induced on the surface of the TI. Addit ionally, the magnetic\ndistribution of Eu has a magnetic domain, which is spatially dependent o n the position on\nthe surface. Therefore, when we move the magnetic domain by usin g an applied magnetic\nfield, the charge density and the spin current are also triggered on ly around the magnetic\ndomain. Recently, magnetic insulator with noncoplanar spin structu re has been reported.\nFor example, magnetoelectric insulator Cu 2OSeO3has spatially dependent spin structure,\nand is called skyrmion, which is topologically protected magnetic spin vo rtex-like object51,52.\nIf one can prepare the vortex-like spin structure deposited on th e surface of the TI and can\ntrigger the magnetization dynamics of the skyrmion, we expect tha t the charge and spin\n28currents are driven by the magnetization dynamics of the skyrmion . Moreover, the spatial\ndistributions of the charge density and the magnitude of the spin cu rrent depends on the\nposition of the skyrmion, because the spatial derivative of the loca lized spin depends on the\npositions in the skyrmion. We comment that the induced spin and char ge currents could be\nindependent of the polarity of the skyrmion, since these current a re triggered by the in-plane\ncomponent of the localized spin [see Eqs. (31) and (42)-(43)]. Then , we expect that in the\nMIs deposited on the surface of the TI, the magnetization dynamic s induces not only the\nlocal spin-charge generation and transport, but also the diffusive one. Our obtained results\nwill enable the applications of TI nanomembrane in spintronics devices .\nACKNOWLEDGEMENTS\nThe authors would like to thank A. A. Golubov, A. Dutt, and A. Yamak aga for valu-\nable discussions. This work was supported by Grants-in-Aid for You ng Scientists (B) (No.\n22740222 and No. 23740236), by Grants-in-Aid for Scientific Rese arch on Innovative Areas\n“Topological Quantum Phenomena” (No. 22103005 and No. 251037 09) from the Ministry\nof Education, Culture, Sports, Science, and Technology, Japan ( MEXT), and by the Core\nResearch for Evolutional Science and Technology (CREST) of the J apan Science. K.T. ac-\nknowledges support from a Grant-in-Aid for Japan Society for the Promotion of Science\n(JSPS) Fellows.\nAppendix A: Derivation of Eqs. (65)-(66)\nWe show details of calculation of coefficient Aq,ΩandBq,Ωof nonlocal localized spin ∝an}b∇acketle{tS∝an}b∇acket∇i}htD.\nThe∝an}b∇acketle{tS∝an}b∇acket∇i}htDis given by\n∝an}b∇acketle{tS∝an}b∇acket∇i}htD(x,t)≡/integraldisplay/integraldisplay\ndt′dx′D(x−x′,t−t′)S(x′,t′) (A1)\nTo substitute S=S(1,0,−i)ei(q·x−Ωt)∝ei(q·x−Ωt)into the above equation, we calculate\n∝an}b∇acketle{tn∝an}b∇acket∇i}htD:\n∝an}b∇acketle{tn∝an}b∇acket∇i}htD=1\nL2/integraldisplay/integraldisplay\ndt′dx′/summationdisplay\nQ,ωei[ω(t−t′)−Q·(x−x′)]ei(q·x′−Ωt′)\nQ2ℓ2+iωτ\n=ei(q·x−ωt)\nq2ℓ2−iΩτ(A2)\n29The resulting ∝an}b∇acketle{tSx∝an}b∇acket∇i}htDis obtained from the real part of S∝an}b∇acketle{tn∝an}b∇acket∇i}htDin the above equation. In the\nsame way, ∝an}b∇acketle{tSz∝an}b∇acket∇i}htDis obtained from the real part of −iS∝an}b∇acketle{tn∝an}b∇acket∇i}htD. Thus, from the above equation,\nthe coefficient of Aq,ΩandBq,Ωare derived as Eqs. (65)-(66).\nAppendix B: Charge conservation\nTo check the validity of the spin current and the charge current we calculate, we use the\ncharge conservation law ∂tρe+∇·j= 0. The charge density ∂tρeis given by\n∂tρe=eνeJsdτ\nL2/bracketleftbigg/summationdisplay\nq,Ωei[Ωt−q·x]iℓΩ2\nq2ℓ2+iΩτ(qySx\nq,Ω−qxSy\nq,Ω)/bracketrightbigg\n, (B1)\nwhereℓ= ˜vFτis the mean-free path. The resulting ∇·jbecomes\n∇xjx=e˜vFνeJsdτ\nL2/summationdisplay\nq,Ωei[Ωt−q·x]Ω/bracketleftbigg/braceleftbigg\n1−1\n2q2ℓ2\nq2ℓ2+iΩτ/bracerightbigg\nqxSy\nq,Ω+1\n2ℓ2q2\nℓ2q2+iΩτqySx\nq,Ω/bracketrightbigg\n,\n∇yjy=−e˜vFνeJsdτ\nL2/summationdisplay\nq,Ωei[Ωt−q·x]Ω/bracketleftbigg/braceleftbigg\n1−1\n2q2ℓ2\nq2ℓ2+iΩτ/bracerightbigg\nqySx\nq,Ω+1\n2ℓ2q2\nℓ2q2+iΩτqxSy\nq,Ω/bracketrightbigg\n,\nand\n∇xjx+∇yjy=e˜vFνeJsdτ\nL2/summationdisplay\nq,Ωei[Ωt−q·x]Ω/bracketleftbigg/braceleftbigg\n1−q2ℓ2\nq2ℓ2+iΩτ/bracerightbigg\n(qxSy\nq,Ω−qySx\nq,Ω)/bracketrightbigg\n.\n=eνeJsdτ\nL2/summationdisplay\nq,Ωei[Ωt−q·x]/bracketleftbiggiΩ2˜vFτ\nq2ℓ2+iΩτ(qxSy\nq,Ω−qySx\nq,Ω)/bracketrightbigg\n=−∂tρe.(B2)\nTherefore, the ρeandjsatisfy∂tρe+∇·j= 0.\n1Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, Spin pumping and magnetization dynamics\nin metallic multilayers , Phys. Rev. Lett. 88, 117601 (2002).\n2J.-I. Ohe, A. Takeuchi, and G. Tatara, Charge Current Driven by Spin Dynamics in Disordered\nRashba Spin-Orbit System , Phys. Rev. Lett. 99, 266603 (2007).\n3E.Saitoh, M. Ueda, H. Miyajima, andG. Tatara, Conversion of spin current into charge current\nat room temperature: Inverse spin-Hall effect , Appl. Phys. Lett. 88, 182509 (2006).\n4A. Stern, Berry’s phase, motive forces, and mesoscopic conductivity , Phys. Rev. Lett. 681022\n(1992).\n305A. Takeuchi and G. Tatara, Charge and Spin Currents Generated by Dynamical Spins , J. Phys.\nSoc. Jpn. 77, 074701 (2008).\n6M. Z. Hasan and C. L. Kane, Colloquium: Topological insulators , Rev. Mod. Phys. 82, 3045\n(2010).\n7X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors , Rev. Mod. Phys. 83,\n1057 (2011).\n8Y. Ando, Topological Insulator Materials , J. Phys. Soc. Jpn. 82, 102001 (2013).\n9L. Fu and C. L. Kane, Superconducting Proximity Effect and Majorana Fermions at t he Surface\nof a Topological Insulator , Phys. Rev. Lett. 100, 096407 (2008).\n10A. R. Akhmerov, J. Nilsson, and C. W. J. Beenakker, Electrically Detected Interferometry of\nMajorana Fermions in a Topological Insulator , Phys. Rev. Lett. 102, 216404 (2009).\n11Y. Tanaka, T. Yokoyama and N. Nagaosa, Manipulation of the Majorana Fermion, Andreev\nReflection, and Josephson Current on Topological Insulators , Phys. Rev. Lett. 103, 107002\n(2009).\n12X.-L. Qi, T. L. Hughes, and S.-C. Zhang, Topological field theory of time-reversal invariant\ninsulators , Phys. Rev. B 78, 195424 (2008).\n13K.NomuraandN. Nagaosa, Electric charging of magnetic textures on the surface of a to pological\ninsulator , Phys. Rev. B 82, 161401 (2010).\n14H. T. Ueda, A. Takeuchi, G. Tatara, and T. Yokoyama, Topological charge pumping effect by the\nmagnetization dynamics on the surface of three-dimensiona l topological insulators , Phys. Rev.\nB85, 115110 (2012).\n15I. Garate and M. Franz, Magnetoelectric response of the time-reversal invariant h elical metal ,\nPhys. Rev. B 81, 172408 (2010).\n16T. Yokoyama, Y. Tanaka, and N. Nagaosa, Anomalous magnetoresistance of a two-dimensional\nferromagnet/ferromagnet junction on the surface of a topol ogical insulator , Phys. Rev. B, 81,\n121401(R) (2010).\n17S. Mondal, D. Sen, K. Sengupta, and R. Shankar, Magnetotransport of Dirac fermions on the\nsurface of a topological insulator , Phys. Rev. B. 82, 045120 (2010).\n18B. D. Kong, Y. G. Semenov, C. M. Krowne, and K. W. Kim, Unusual magnetoresistance in a\ntopological insulator with a single ferromagnetic barrier , App. Phys. Lett. 98, 243112 (2011).\n19D. Culcer, Transport in three-dimensional topological insulators: Th eory and experiment , Phys-\n31ica E,44, 860, (2012).\n20M. J. Ma, M. B. A. Jalil, S. G. Tan, Y. Li, and Z. B. Siu, Spin current generator based on\ntopological insulator coupled to ferromagnetic insulator s, AIP Advances 2, 032162 (2012)\n21A. A. Burkov and D. G. Hawthorn, Spin and Charge Transport on the Surface of a Topological\nInsulator , Phys. Rev. Lett. 105, 066802 (2010).\n22P. Schwab, R. Raimondi, and C. Gorini, Spin-charge locking and tunneling into a helical metal ,\nEPL (Europhysics Letters), 93, 67004 (2011).\n23K. Taguchi, T. Yokoyama and Y. Tanaka, Giant magnetoresistance in the junction of two\nferromagnets on the surface of diffusive topological insula tors, Phys. Rev. B 89, 085407 (2014).\n24T. Yokoyama, J. Zang, and N. Nagaosa, Theoretical study of the dynamics of magnetization on\nthe topological surface , Phys. Rev. B 81, 241410 (2010).\n25A. Sakai and H. Kohno, Spin torques and charge transport on the surface of topologi cal insulator ,\nPhys. Rev. B 89, 165307 (2014).\n26A. R. Mellnik et al., Spin-transfer torque generated by a topological insulator , Nature (London)\n511, 449 (2014).\n27P. Deorani, J. Son, K. Banerjee, N. Koirala, M. Brahlek, S. Oh , and H. Yang, Observation of\ninverse spin Hall effect in bismuth selenide , Phys. Rev. B 90, 094403 (2014).\n28Y. Shiomi, K. Nomura, Y. Kajiwara, K. Eto, M. Novak, Kouji Seg awa, Yoichi Ando, and E.\nSaitoh,Spin-Electricity Conversion Induced by Spin Injection into To pological Insulators , Phys.\nRev. Lett. 113196601 (2014).\n29M. Jamali, J.-S. Lee, Y. Lv, Z. Zhao, N. Samarth, and J.-P. Wan g,Room Temperature Spin\nPumping in Topological Insulator Bi2Se3, arXiv:1407.7940\n30X. Liu and J. Sinova, Reading Charge Transport from the Spin Dynamics on the Surfac e of a\nTopological Insulator , Phys. Rev. Lett. 111, 166801 (2013).\n31M. H. Fischer, A. Vaezi, A. Manchon, and E.-Ah Kim, Large Spin Torque in Topological Insu-\nlator/Ferromagnetic Metal Bilayers , arXiv:1305.1328.\n32T. Yokoyama and Y. Tserkovnyak, Spin diffusion and magnetoresistance in\nferromagnet/topological-insulator junctions , Phys. Rev. B 89, 035408 (2014).\n33G. D. Mahan, Many-Particle Physics , 3rd ed. (Springer, 2000), p.514.\n34J. Fujimoto, A. Sakai, H. Kohno, Ultraviolet divergence and Ward-Takahashi identity in a tw o-\ndimensional Dirac electron system with short-range impuri ties, Phys. Rev. B 87, 085437 (2013).\n3235Using/planckover2pi1\n2τ=1\n2πνeniu2\n0,νe=ǫF\n2π/planckover2pi12˜v2\nF, and ˜vF/vF= (1+ξ)−1, wecanobtaintherelation: ξ(1+ξ)2=\n/planckover2pi1\n4πǫFτ. Because of/planckover2pi1\nǫFτ≪1,ξ 0additional frequencies appear, and the\ncorresponding oscillations extend to longer times while th e\noriginal peaks become smaller and some split into two.\nForΩ = 0 , the behavior changes from exponential to\nalgebraic.\nFig. 2 shows the correlations for typical experimental\nparameters and different SO coupling constants Ω = 0 .\nAs also seen in equation 3, ∆oscillates with the Larmor\nfrequency differences ωi−ωjforΩ = 0 . In GaAs, the\nthreeγiare approximately equidistant. Therefore, the\nrelative orientations of the three contributions to ∆B⊥\nnuc\nand thus ∆return to nearly the same value after one\nperiodTij= 2π(ωi−ωj)−1of the slower relative pre-\ncession,T75−69≈T71−69. Hence, large revivals in the\ncorrelations occur at this value of δt. The smaller peaks\natδt=T75−71result from a partial realignment of these\ntwo species only. The envelope decay of the correlations\non a time scale of approximately 40 µs arises from nuclear\nspin dephasing.\nAsΩis increased, oscillations at the bare frequencies\nωiextending to larger δtstart to appear as well. In addi-\ntion, the large peak around δt= 17µssplits. Observing\nthese features would represent qualitative evidence for\nSO coupling. At the same time, the magnitude of the\nvariation in the correlations is reduced.\nWe now turn to proof of principle experimental results,\nobtained from the same DQD sample as used in [19]. In\nall measurements, the sweep range was constrained to\nkeepJ(ε)≫Bnucto prevent mixing of S and T 0. As\nBext≫∆, the nonlinear sweep and the finite sweep range4\ndo not introduce a large error. A measured background\nreflecting the direct coupling of the gate pulses to the\ncharge sensor used for single shot readout was subtracted\nfrom each dataset.\nTo verify the expected behavior of flip the probability,\nwe measured its dependence on sweep rate and magnetic\nfield, as shown in Fig. 3(a). The scaling factor used\nto convert the readout signal to triplet probability was\nchosen in accordance with the contrast of the (1,1)-(0,2)\ncharge transition, taking relaxation during the readout\nstage into account. The scaling of the x-axis is motivated\nby the fact that in the experiment the sweeps result in a\nlinear time dependence of εrather than J(ε). To com-\npute the sweep rate α=dJ/dt=dJ/dε·dε/dt, we employ the\nphenomenological relation J(ε) =J0exp/parenleftBig\n−ε\nε0/parenrightBig\n[21]. At\nthe S-T+ transition, J(ε) =Bext, so thatα=Bext1\nε0∆ε\n∆t,\nwhere∆tis time taken to sweep detuning by ∆ε. Hence,\nPSis expected to be a function of ∆t/Bext, independent\nofBext. Indeed, the curves for different Bextin Fig. 3(a)\ncollapse reasonably well onto each other. The deviations\nfor slow sweeps may be a result of relaxation. The solid\nline was obtained from Eq. 4 using Ω = 0 and a value\nof∆2/αthat is a factor 6 smaller than expected based\non measurements of T∗\n2, from which N= 2.3·106can be\nextracted [19]. This adjustment was required to obtain\nadequate agreement with the model, which contains no\nfurther free parameters. As can be seen from the rela-\ntively large slope at large ∆t, the algebraic dependence\nobtained from nuclear averaging in equation 4 gives a bet-\nter overall fit than the exponential LZ formula (dashed\nline), even if adjusting the scaling factor of the latter.\nGuided by these results, we fixed ∆tat 0.3µs for the\ncorrelation measurements discussed next. Initialization –\nLZ-sweep–readout cycles as shown in Fig. 1(b) were re-\npeated at a rate of 0.5 MHz. Initialization was carried\nout via a standard procedure through electron exchange\nwith the leads [2]. Readout was accomplished via RF re-\nflectometry of a quantum point contact [15]. Instead of\nsingle shot discrimination, we computed the correlations\ndirectly from the readout signal, which was averaged over\n1µsin each cycle. Assuming the readout signal is the\nsum of a contribution from the qubit measurement and\nnoise uncorrelated with the qubit state, this procedure is\nequivalent to correlating single shot readout values apart\nfrom additional noise. Due to the fixed sampling rate,\nthe time resolution was limited to 2 µs.\nThe data shown in Fig. 3(b) clearly shows the expected\noscillations. For fitting our model to the experiment,\nonly an offset of the QPC signal was adjusted individ-\nually for each data set. The y-scale was set analogous\nto Fig. 3(a). Obtaining good agreement of the mag-\nnitude required ∆2/αto be chosen a factor 18 smaller\nthan expected for N= 2.3·106unit cells and experi-\nmental sweep rates. The δt-dependence is unaffected by\nthis adjustment. Nuclear spin Larmor frequencies were \nΔt / Bext ( μs / T) \n0 20 40 60 80 10000.050.10.150.20.250.3 Bext = 40 mT \nBext = 35 mT \nBext = 30 mT \nBext = 25 mT \nBext = 20 mT \nBext = 15 mT 0 10 20 30 40 50 60 70 80 00.250.50.751\nBext = 75 mT Bext = 15 mT \nδt (μs) PSPS(t)PS(t+δt)(a)\n(b) 1 - \nFigure 3. (a) Triplet probabilities as a function of sweep-t ime\n∆tfrom 0µs to 1µs with a constant ∆εfor different Bext.\nThe scaling of the time axis with magnetic field leads to a\nreasonable collapse of the curves. Data in color. The black\nlines shows the LZ-model (dashed) with a fixed transition\nmatrix element, and the nuclear spin averaged LZ behavior\n(solid). With N= 2.3·106fixed,∆2/αrequires a scaling\nfactor of 6 to match experiment. (b) Correlation of triplet\nprobabilities for different Bext, data in black and model in\nred. Except for the magnitude, the correlations behave as\nexpected as a function of Bext.\ntaken from [22]. δBlocdetermines the time scale of the\noverall decay and was fixed to 0.33 mT, consistent with\nearlier measurements [19]. As expected for the low mag-\nnetic fields considered here, the effect of SO coupling is\nnegligible. At higher fields, contrast is lost faster than\nexpected from our model, but the time resolution used\nhere also becomes too coarse to resolve the Larmor pre-\ncession.\nThe experiments thus confirm the overall picture ob-\ntained from our model and the usefulness of single shot\ncorrelation measurements. The significant quantitative\ndiscrepancy in ∆2/αis not understood, but may at least\npartly arise from noise and non-uniformity of the sweep5\nresulting from the discrete 1 ns sampling interval of the\nwave form generator. More detailed measurements would\nbe required to explore these effects. By varying the de-\nlay between subsequent pulses instead of using a fixed\nrepetition rate, the time resolution could be improved\nsignificantly. This would enable an extension to larger\nmagnetic fields where SO coupling might become impor-\ntant.\nIn conclusion, we have shown that correlation measure-\nments of Landau-Zener sweeps through the S-T+ transi-\ntion reveal nuclear spin precession and dephasing. These\ndynamics are seen in an experimental implementation of\nthe scheme. They imply that the perpendicular compo-\nnent of the Overhauser field cannot be controlled as well\nas thez-component. Our model indicates that the same\nmeasurement technique should be able to qualitatively\nreveal SO coupling at higher magnetic fields.\nThis work was supported by the Alfried Krupp von\nBohlen und Halbach Foundation and DFG under Grant\nNo. BL 1197/2-1. We would like to thank Amir Yacoby,\nMark Rudner and Izhar Neder for discussions.\n∗Present address: QuTech and Kavli Institute of\nNanoscience, Delft University of Technology, 2600 GA\nDelft, The Netherlands\n[1] D. Loss and D. P. DiVincenzo,\nPhys. Rev. A 57, 120 (1998).\n[2] J. R. Petta, H. Lu, and A. C. Gossard,\nScience 327, 669 (2010).\n[3] H. Ribeiro, J. R. Petta, and G. Burkard,\nPhys. Rev. B 87, 235318 (2013).\n[4] H. Ribeiro, G. Burkard, J. R. Petta, H. Lu, and A. C.\nGossard, Phys. Rev. Lett. 110, 086804 (2013).\n[5] D. Stepanenko, M. Rudner, B. I. Halperin, and D. Loss,Phys. Rev. B 85, 075416 (2012).\n[6] H. Bluhm, S. Foletti, D. Mahalu, V. Umansky, and\nA. Yacoby, Phys. Rev. Lett. 105, 216803 (2010).\n[7] S. Foletti, H. Bluhm, D. Mahalu, V. Umansky, and\nA. Yacoby, Nature Physics 5, 903 (2009).\n[8] C. Kloeffel, P. A. Dalgarno, B. Urbaszek, B. D. Gerardot,\nD. Brunner, P. M. Petroff, D. Loss, and R. J. Warburton,\nPhys. Rev. Lett. 106, 046802 (2011).\n[9] C. Latta, A. Hoegele, Y. Zhao, A. N. Vamivakas,\nP. Maletinsky, M. Kroner, J. Dreiser, I. Carusotto,\nA. Badolato, D. Schuh, W. Wegscheider, M. Atature,\nand A. Imamoglu, Nature Physics 5, 758 (2009).\n[10] T. D. Ladd, D. Press, K. De Greve, P. L. McMahon,\nB. Friess, C. Schneider, M. Kamp, S. Höfling, A. Forchel,\nand Y. Yamamoto, Phys. Rev. Lett. 105, 107401 (2010).\n[11] M. S. Rudner, I. Neder, L. S. Levitov, and B. I. Halperin,\nPhys. Rev. B 82, 041311 (2010).\n[12] A. Brataas and E. I. Rashba,\nPhys. Rev. B 84, 045301 (2011).\n[13] A. Brataas and E. I. Rashba,\nPhys. Rev. Lett. 109, 236803 (2012).\n[14] I. Neder, M. S. Rudner, and B. I. Halperin,\nPhys. Rev. B 89, 085403 (2014).\n[15] C. Barthel, D. J. Reilly, C. M. Marcus, M. P. Hanson,\nand A. C. Gossard, Phys. Rev. Lett. 103, 160503 (2009).\n[16] T. Fink and H. Bluhm,\nPhys. Rev. Lett. 110, 010403 (2013).\n[17] L. Landau, Physics of the Soviet Union 2 , 46 (1932).\n[18] C. Zener, Proc. Roy. Soc. London, Ser. A 137, 696 (1932).\n[19] H. Bluhm, S. Foletti, I. Neder, M. Rudner,\nD. Mahalu, V. Umansky, and A. Yacoby,\nNature Physics 7, 109 (2011).\n[20] I. Neder, M. S. Rudner, H. Bluhm, S. Fo-\nletti, B. I. Halperin, and A. Yacoby,\nPhys. Rev. B 84, 035441 (2011).\n[21] O. E. Dial, M. D. Shulman, S. P. Harvey,\nH. Bluhm, V. Umansky, and A. Yacoby,\nPhys. Rev. Lett. 110, 146804 (2013).\n[22] D. Paget, G. Lampel, B. Sapoval, and V. I. Safarov,\nPhys. Rev. B 15, 5780 (1977)." }, { "title": "0801.4132v2.Spin_dynamics_of_two_dimensional_electrons_with_Rashba_spin_orbit_coupling_and_electron_electron_interactions.pdf", "content": "arXiv:0801.4132v2 [cond-mat.mes-hall] 3 Nov 2008Spin dynamics of two-dimensional electrons with Rashba spi n-orbit coupling and\nelectron-electron interactions\nYuan Li1,2and You-Quan Li1\n1Department of Physics, Zhejiang University, Hangzhou 310027, P . R. China\n2Institute of Materials Physics, Hangzhou Dianzi University, Hangz hou 310018, P. R. China\n(Dated: October 29, 2018)\nWe study the spin dynamics of two dimensional electron gases (2DEGs) with Rashba spin-orbit\ncoupling by taking account of electron-electron interacti ons. The diffusion equations for charge\nand spin densities are derived by making use of the path-inte gral approach and the quasiclassical\nGreen’s function. Analyzing the effect of the interactions, we show that the spin-relaxation time\ncan be enhanced by the electron-electron interaction in the ballistic regime.\nPACS numbers: 72.25.Rb, 71.70.Ej, 71.10.Ca\nI. INTRODUCTION\nSpin based electronics [1] or spintronics [2] has been\nan active research area in the past decade. The effort\nfor effectively manipulating electron spin by means of an\napplied electric field [3, 4, 5] is an important issue there.\nThe system with spin-orbit (SO) couplings makes those\nefforts possible and thus brings great interests from both\nacademicandpracticalaspectsrecently. Thus,itisessen-\ntial to study the spin relaxation for further development\nof spintronics.\nThere are four main mechanisms of spin relaxation in\nsemiconductor systems [2, 6, 7, 8, 9, 10]. In Elliott-Yafet\nmechanism, the spin-orbit coupling induces a mixing of\nwave functions for valence-band states and conduction-\nband states. The mixing that results in the spin relax-\nation ofelectrons is due to the scattering by impurities or\nphonons. The Elliott-Yafet mechanism operates in semi-\nconductors with and without a center of inversion sym-\nmetry, while it is most prominent in the centrosymmetric\nones (such as silicon). The Bir-Aronov-Pikus mechanism\nis applicable for p-doped semiconductors in which the\nelectron spin flipping is induced by exchange interaction\nwith holes. The hyperfine interaction provides another\nimportant mechanism [11] for ensemble spin dephasing\nand single spin decoherence of localized electrons. The\nD’yakonov-Perel mechanism depicts that electrons can\nfeel an effective random magnetic field arising from the\nspin-orbit coupling in systems with inversion asymme-\ntry such that spin relaxation occurs. This mechanics\ncan interpret the spin dephasing in crystals without in-\nversion center and is particularly applicable for n-type\nsamples. For two dimensional n-type semiconductor sys-\ntems without inversion symmetry, the D’yakonov-Perel\nmechanism is believed to be most important in wide\nranges of carrier temperature and concentration. Under\ncertain conditions, the Elliott-Yafet mechanism may af-\nfect spin dynamics of two-dimensional electrons in these\nsystems. The Bir-Aronov-Pikus mechanism is impor-\ntant forp-type semiconductor systems and the hyperfine-\ninteraction mechanism dominates for localized electrons.\nMost studies of spin relaxation in semiconductors havefocused on impurity (somewhat less phonon) mediated\nspin flips while neglecting the effect of electron-electron\ninteractions for a long time. It has been noticed re-\ncentlythatelectron-electroninteractionsplaycertainrole\nin spin relaxation and dephasing in semiconductor sys-\ntems. The electron-electron interaction is known to play\na crucial role in determining the transport and thermo-\ndynamic properties near the metal-insulatortransition in\ntwo-dimensionalelectronsystems[12], whichissuspected\nto affect the spin relaxation for the spin susceptibility\nbehaving critically when the metal-insulator transition\noccur [13, 14, 15]. There are several experimental and\ntheoretical studies on the effect of electron-electron in-\nteractions on spin relaxation. The electron-electron scat-\ntering results in additional momentum relaxation which\ninduces spin dephasing of electrons through the motional\nnarrowing of the D’yakonov-Perel type [16] as measured\ninn-GaAs/AlGaAsquantumwells[17, 18]. Theelectron-\nelectron scattering effect on the spin dephasing has been\nconsidered[19] in a magnetic field, and a momentum de-\npendent effective random magnetic field induced by the\nelectron-electron exchange interaction can lead to spin\ndephasing of electrons [20, 21, 22]. It is also observed\nthat the spin relaxation caused by the D’yakonov-Perel\nmechanism gives considerably different rates depending\non the technique employed [23].\nHowever, as we are aware, the explicit form of the\ndiffusion equation for two-dimensional electron gases\n(2DEGs) with spin-orbit couplings has not been derived\nby taking account of electron-electron interactions. It is\nthus obligatory to develop the explicit form of the dif-\nfusion equation to study the spin dynamics for 2DEGs\nwith spin-orbit couplings as well as electron-electron in-\nteractions. In this paper, we focus attention on the\nD’yakonov-Perel spin-relaxation mechanism. We inves-\ntigate the spin dynamics of electrons in two-dimensional\nn-type semiconductor systems with electron-electron in-\nteractions and Rashba spin-orbit coupling.\nThe paper is organized as follows. In Sec. II, we\ntake account of the electron-electron interaction for the\n2DEGs with the Rashba spin-orbit coupling. Applying\nthe path integral formulation, we decoupled the interac-2\nFIG. 1: The Keldysh contour.\ntion in terms of an auxiliary Bose field. In Sec. III, we\nemploy the quasiclassical Green’s function to investigate\nthe spin dynamics of electrons. In Sec. IV, the diffusion\nequations for spin and charge densities as well as the ex-\nplicit expression of spin-relaxation time are derived. A\nsummary is given in Sec. V and some complicated for-\nmulae are given in the Appendix.\nII. AUXILIARY FIELDS DESCRIBING THE\nELECTRON-ELECTRON INTERACTION\nTaking the electron-electron interaction into account,\nwe study the spin dynamics of electrons in two-\ndimensionalsystemswith structureinversionasymmetry.\nAs the Fourier transform of the Coulomb repulsion be-\ntween electrons reads V(q) = 2πe2/|q|, the Hamiltonian\nof such a system is given by\nˆH=/integraldisplay/braceleftig/summationdisplay\nλ,λ′ˆψλ†(r)/bracketleftbig/parenleftbig\n−/planckover2pi12\n2m∇2+U(r)−µ/parenrightbig\nδλ,λ′\n+b·/vector σλλ′/bracketrightbigˆψλ′(r)/bracerightig\nd2r (1)\n+1\nA/summationdisplay\nq/negationslash=0πe2\n|q|ˆρ(q)ˆρ(−q),\nwhereˆψλ†(r) andˆψλ(r) representthefieldoperatorswith\nλ=↑,↓labelling the spin state of the electron, ˆ ρ(q)\nrepresents the Fourier transform of the density operator\nˆρ(r) =/summationtext\nλˆψλ†(r)ˆψλ(r) and/vector σ= (σx,σy,σz) the Pauli\nmatrices in spin space, U(r) a random disorder poten-\ntial andµthe chemical potential. The other notions in\nEq. (1) are A=L2withLreferring to the size of the\nsample and b=αp×ezwithαreferring to the Rashba\nspin-orbit coupling strength. In holonomyrepresentation\n(or the called coherent state representation), the Green’s\nfunction can be expressed as a functional integral over\ntheGrassmannfields ψλand¯ψλthatreflectthefermionic\nnature of electrons,\nGλλ′(r,t;r′,t′) =∝an}b∇acketle{tψλ(r,t)¯ψλ′(r′,t′)∝an}b∇acket∇i}ht\n=/integraldisplay\nD¯ψDψψλ(r,t)¯ψλ′(r′,t′)e−iS[ψ,¯ψ]\n/integraldisplay\nDψD¯ψe−iS[ψ,¯ψ].(2)Here we adopted the unit /planckover2pi1= 1 and the simplified no-\ntationDψ=Dψ↑Dψ↓. The action S[ψ,¯ψ] in the above\nequation is given by\nS[ψ,¯ψ] =/integraldisplay\ndt/braceleftig/integraldisplay\nd2r/summationdisplay\nλλ′¯ψλ(r,t)Wλλ′ψλ′(r,t)\n+1\nA/summationdisplay\nq/negationslash=0πe2\n|q|/parenleftbig\nρ(q,t)ρ(−q,t)/parenrightbig/bracerightig\n,(3)\nwhereWλλ′=/parenleftbig\n−i∂/∂t−∇2\nr/2m+U(r)−µ/parenrightbig\nδλλ′+b·/vector σλλ′.\nWe divide the fermionic field ψλ(r,t) into two compo-\nnentsψλ\n1(r,t) andψλ\n2(r,t) which reside, respectively, on\nthe upper and lower branches of the Keldysh time con-\ntour shown in Fig. (1). Hence, the second line of Eq. (3),\nwhich refers to the interaction part, can be written as\nSint[ψ1,¯ψ1]−Sint[ψ2,¯ψ2] with\nSint[ψi,¯ψi] =/integraldisplay\ndt/summationdisplay\nq/negationslash=0πe2\nA|q|/parenleftbig\nρi(q,t)ρi(−q,t)/parenrightbig\n,\nwherei= 1,2. With the help of two auxiliary bosonic\nfields˜φi(r,t), we can decouple those two terms rele-\nvant to electron-electron interactions via the Hubbard-\nStratonovich transformation [24], namely,\nexp/bracketleftbig\n−i/integraldisplay\ndt/summationdisplay\nq/negationslash=0πe2\nA|q|ρi(q,t)ρi(−q,t)/bracketrightbig\n=/integraldisplay\nD˜φi(q,t)exp/bracketleftbig\ni/integraldisplay\ndt/summationdisplay\nq/negationslash=0|q|\n4π˜φi(q,t)˜φi(−q,t)/bracketrightbig\n×exp/bracketleftig\ni/integraldisplay\ndte\n2√\nA/summationdisplay\nq/negationslash=0/braceleftbig˜φi(q,t)ρi(−q,t)\n+ρi(q,t)˜φi(−q,t)/bracerightbig/bracketrightig\n.\nThen we can write the Green’s function as follows,\nˆGλλ′(r,t;r′,t′)\n=/integraldisplay\nD¯ΨDΨDΦΨλ(r,t)¯Ψλ′(r′,t′)e−iS[Ψ,¯Ψ,Φ]\n/integraldisplay\nDΨD¯ΨDΦe−iS[Ψ,¯Ψ,Φ],\n(4)\nwhere the action in real space is given by3\nS[Ψ,¯Ψ,Φ] =/integraldisplay\ndtd2r/braceleftig/summationdisplay\nλλ′¯Ψλ(r,t)/bracketleftbig\nWλλ′σ3−e˜φi˜γiδλλ′/bracketrightbig\nΨλ′(r,t)/bracerightig\n+/integraldisplay\ndt/integraldisplay\nd2rd2r′/braceleftig\nΦT(r,t)−e2\n2V−1\n0(r−r′)σ3Φ(r′,t)/bracerightig\n, (5)\nin which the Pauli matrix σ3= diag(1,−1) is defined on\nthe Keldysh space, and V−1\n0is defined via the following\nrelation\n/integraldisplay\nd2r1V0(r−r1)V−1\n0(r1−r′) =δ(r−r′).\nThe other notions appeared in Eq. (5) are fermionic dou-\nblet Ψ, bosonic doublet Φ, and vertex matrices ˜ γi, they\nare defined as\nΨλ=/parenleftbigg\nψλ\n1\nψλ\n2/parenrightbigg\n,Φ =/parenleftbigg˜φ1\n˜φ2/parenrightbigg\n,\n˜γ1=/parenleftbigg\n1 0\n0 0/parenrightbigg\n,˜γ2=/parenleftbigg\n0 0\n0−1/parenrightbigg\n.\nFor calculation convenience, one can introduce a parti-\ntion function for the coupling between the fermionic and\nbosonic doublets,\nZ[Φ] =∝an}b∇acketle{tTCe−iSR[Φ,Ψ]∝an}b∇acket∇i}htΨ,\nSR[Φ,Ψ] =/integraldisplay\ndt d2r/braceleftig/summationdisplay\nλ¯Ψλ/parenleftbig\n−e˜φi˜γi/parenrightbig\nΨλ/bracerightig\n,\nwhere T Cstands for time ordering along the contour C\nand∝an}b∇acketle{t ··· ∝an}b∇acket∇i}ht Ψmeans functional integration over Ψ field\nwith the action\nS[Ψ] =/integraldisplay\ndtd2r/braceleftig/summationdisplay\nλλ′Wλλ′¯Ψλ(r,t)σ3Ψλ′(r,t)/bracerightig\n.\nThen the Green’s function in Eq. (4) can be formally\nexpressed as a functional integration over the bosonic\nfields,\nˆGλλ′(r,t;r′,t′) =N/integraldisplay\nDΦˆGλλ′(r,t;r′,t′|Φ)\n×exp/braceleftbig\n−iSe[Φ]/bracerightbig\n, (6)\nwhere the normalization coefficient is denoted by Nand\nthe actionSe[Φ] is defined by\nSe[Φ] =ilnZ[Φ]+/integraldisplay\ndt d2rd2r′\n×/braceleftbig\nΦT(r,t)−e2\n2V−1\n0(r−r′)σ3Φ(r′,t)/bracerightbig\n,(7)\nand the kernel ˆG(r,t;r′,t′|Φ) is given by\nˆGλλ′(r,t;r′,t′|Φ)\n=1\nZ[Φ]∝an}b∇acketle{tTCΨλ(r,t)¯Ψλ′(r′,t′)e−iSR[Φ,Ψ]∝an}b∇acket∇i}htΨ.(8)We can averagethe Green’s function ˆGλλ′(r,t;r′,t′) over\ndisorder as follows [25]\n∝an}b∇acketle{tˆGλλ′(r,t;r′,t′)∝an}b∇acket∇i}htdis=N/integraldisplay\nDΦ∝an}b∇acketle{tˆGλλ′(r,t;r′,t′|Φ)∝an}b∇acket∇i}htdis\n×exp/braceleftbig\n−i∝an}b∇acketle{tSe[Φ]∝an}b∇acket∇i}htdis/bracerightbig\n,(9)\nwhere∝an}b∇acketle{t···∝an}b∇acket∇i}htdisrefers to the average over disorder. The\nrandom disorder potential U(r) is assumed to be charac-\nterized by a correlation function\n∝an}b∇acketle{tU(r)U(r′)∝an}b∇acket∇i}htdis=1\n2πντδ(r−r′),\nwhereν=m/π/planckover2pi12stands for the density of states. The\naverage of the Green’s function over disorder introduces\nthe elastic scattering time τwhich is relevant to the ran-\ndom disorder. We neglect correlationsbetween the meso-\nscopic fluctuations of ∝an}b∇acketle{tSe[Φ]∝an}b∇acket∇i}htdisand the fermionic opera-\ntors in Eq. (8) so that the averageofthe Green’s function\nˆGλλ′(r,t;r′,t′|Φ) can be separated from the bosonic ac-\ntionSe[Φ]. This approximation is valid since the meso-\nscopic fluctuation is smaller than average quantities.\nAfter averaging over disorder, we rotate the Keldysh\nbasesˆG→Lσ3ˆGL†through a unitary matrix L\nL=1√\n2/parenleftbigg\n1−1\n1 1/parenrightbigg\n,\nso that the Green’s function takes the following shape,\nˆG(r,t;r′,t′|Φ)\n=/parenleftbigg\nGR(r,t;r′,t′|Φ)GK(r,t;r′,t′|Φ)\nGZ(r,t;r′,t′|Φ)GA(r,t;r′,t′|Φ)/parenrightbigg\n.(10)\nNote that the Green’s function ˆG(r,t;r′,t′|Φ) is a 2 ×2\nmatrix defined in the Keldysh space, of which the matrix\nentities are again 2 ×2 matrices defined in spin space.\nThe bosonic fields after rotation take the following two\ncomponents φ1=e\n2(˜φ1+˜φ2) andφ2=e\n2(˜φ1−˜φ2) that\nreside on the upper and lower branches of the contour\nC, respectively. Then the corresponding vertex matrices\nturn toγ1(2)=L(˜γ1±˜γ2)σ3L†, namely,\nγ1=/parenleftbigg\n1 0\n0 1/parenrightbigg\n, γ2=/parenleftbigg\n0 1\n1 0/parenrightbigg\n,\nand the interaction term ¯Ψλ(−e˜φi˜γi)Ψλbecomes\n−¯Ψλ/parenleftbigg\nφ1φ2\nφ2φ1/parenrightbigg\nΨλ. Asγ1andγ2constitute a repre-\nsentation of Z2group, the interaction can be regarded as\nthe coupling between Fermi field and Z2Bose field.4\nThese Bose fields define the following propagators:\nDK(r1,r2;t1,t2) =−2i∝an}b∇acketle{tφ1(r1,t1)φ1(r2,t2)∝an}b∇acket∇i}ht,\nDR(r1,r2;t1,t2) =−2i∝an}b∇acketle{tφ1(r1,t1)φ2(r2,t2)∝an}b∇acket∇i}ht,\nDA(r1,r2;t1,t2) =−2i∝an}b∇acketle{tφ2(r1,t1)φ1(r2,t2)∝an}b∇acket∇i}ht,\n∝an}b∇acketle{tφ2(r1,t1)φ2(r2,t2)∝an}b∇acket∇i}ht= 0. (11)\nOne can show that those propagators obey the Dysonequations in the saddle point approximation, namely,\nˆD(x1,x2) =ˆD0+/integraldisplay\ndx3dx4ˆD0(x1,x3)ˆΠ(x3,x4)ˆD(x3,x2),\n(12)\nwith notations x≡(r,t),DR\n0(q) =DA\n0(q) =−2πe2/|q|,\nand\nˆD=/parenleftbigg\nDRDK\n0DA/parenrightbigg\n,ˆD0=/parenleftbigg\nDR\n00\n0DA\n0/parenrightbigg\n,ˆΠ =/parenleftbigg\nΠRΠK\n0 ΠA/parenrightbigg\n,\nΠR(x1,x2) = ΠA(x2,x1) =δTrsGK(r1;t1,t1|Φ)\n−2iδφ1(r2,t2),\nΠK(x1,x2) =δTrs/bracketleftbig\nGK(r1;t1,t1|Φ)+GZ(r1;t1,t1|Φ)/bracketrightbig\n−2i δφ2(r2,t2),\n(13)\nwhere Tr sstands for the trace in the spin space.\nIII. THE KINETIC EQUATION\nIn previous section, the electron-electron interaction\nhas been decoupled with the help of auxiliary bosonic\nfieldsφ1(2). This means that the influence of the interac-\ntion can be described by a Z2Bose field,\nˆϕ(r,t) =/parenleftbigg\nφ1(r,t)φ2(r,t)\nφ2(r,t)φ1(r,t)/parenrightbigg\n.\nNow we are able to apply the quasiclassicalGreen’s func-\ntion[26, 27, 28]approachtostudythespindynamics. We\nderive the Eilenberger equation from the right-hand and\nleft-hand Dyson equations obeyed by the Green’s func-\ntionˆG(r,t;r′,t′|Φ) in Eq. (10):\n˜∂tˆg+vF·/vector∇ˆg+i/bracketleftbig\nb·/vector σ,ˆg/bracketrightbig\n=ˆg∝an}b∇acketle{tˆg∝an}b∇acket∇i}htn−∝an}b∇acketle{tˆg∝an}b∇acket∇i}htnˆg\n2τ,(14)\nwherevFdenotes the Fermi velocity, τis the elastic scat-\ntering time arising from the adoption of the standard\nself-consistent Born approximation; ∝an}b∇acketle{t···∝an}b∇acket∇i}htnmeans tak-\ning average over the direction of the electron momentum\nn=p/|p| ≡(cosθ,sinθ), and the covariant derivative\nis defined by\n˜∂tˆg=∂t1ˆg+∂t2ˆg+iˆϕ(r,t1)ˆg−iˆgˆϕ(r,t2).(15)\nThe quasiclassical Green’s function in Keldysh and spin\nspaces,\nˆg=/parenleftbigg\ngRgK\ngZgA/parenrightbigg\n, (16)can be derived by integrating the Fourier transform of\nthe Green’s function in Eq. (10) over energy variables,\ni.e.,\nˆg(t1,t2;n,r) =i\nπ/integraldisplay\ndξˆG(t1,t2;p,r),\nˆG(t1,t2;p,r) =/integraldisplay\nd2r′eip·r′ˆG(r1,t1;r2,t2|Φ),(17)\nwhereξ=p2/2m−µ,r′=r1−r2,r= (r1+r2)/2. The\nelectron polarization operators can be obtained in terms\nof Eq. (13) and Eq. (17), i.e.,\nΠR(x1,x2) = ΠA(x2,x1)\n=ν/integraldisplaydθ\n2π/bracketleftbig\nδ(x1−x2)+π δTrsgK(t1,t1;n,r1)\n2δφ1(r2,t2)/bracketrightbig\n,\nΠK(x1,x2)\n=πν/integraldisplaydθ\n2πδTrs/bracketleftbig\ngK(t1,t1;n,r1)+gZ(t1,t1;n,r1)/bracketrightbig\n2δφ2(r2,t2).\n(18)\nSince physical observables are determined by the\nKeldysh component of the quasiclassical Green’s func-\ntion, namely ∝an}b∇acketle{tgK(t1,t2;n,r)∝an}b∇acket∇i}htΦ(here the subscript Φ\nrefers that the functional average [29] is taken over the\nfield Φ), we need to solve this component from the\nEilenberger equation. Decomposing the Green’s func-\ntion in chargeand spin components ∝an}b∇acketle{tgK(t1,t2;n,r)∝an}b∇acket∇i}htΦ=\ngK\n0+gK·/vector σ, one can obtain the charge and spin densities,5\nrespectively,\nρ(r,t) =−1\n4eν/integraldisplay\ndǫ∝an}b∇acketle{tgK\n0(t,ǫ;n,r)∝an}b∇acket∇i}htn,\nS(r,t) =−1\n4ν/integraldisplay\ndǫ∝an}b∇acketle{tgK(t,ǫ;n,r)∝an}b∇acket∇i}htn.(19)\nNow we turn to the kinetic equations for the two inde-\npendent components gKandgZ. For∝an}b∇acketle{tgZ∝an}b∇acket∇i}htΦ= 0 in all\norders of the perturbation theory, we have\ngK=∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ+δgK, gZ=δgZ,\nwhere the fluctuation parts δgimply the effects con-\ntributed by the auxiliary bosonic fields. One can obtain\nfrom Eq. (14) that δgZobeys the following equation,\n(˜∂t+vF·/vector∇)δgZ+i/bracketleftbig\nb·/vector σ,δgZ/bracketrightbig\n−1\nτ/bracketleftbig\nδgZ−∝an}b∇acketle{tδgZ∝an}b∇acket∇i}htn/bracketrightbig\n=−2iφ2(r,t1)δ(t1−t2)Is, (20)\nwhere I sdenotes the unit matrix in spin space. When\nderiving the above equation, we have used the conditions\ngR=δ(t1−t2)Is−gKδgZ/2 andgA=−δ(t1−t2)Is+\nδgZgK/2. Equation (20) gives rise to\nδgZ(t1,t2;n,r) = 2iδ(t1−t2)/integraldisplay\ndr1dt3/integraldisplaydθ′\n2π\n×φ2(r1,t3)Γρ(t3−t1,n′,n;r1,r),\nΓρ(t,n′,n;r1,r2) =/integraldisplaydωd2q\n(2π)3eiq·(r1−r2)−iωt\n×Γρ(n′,n;ω,q), (21)where the diffusion propagator Γ ρis defined by\n(−iω+ivFn·q)Γρ(n,n′;ω,q)+1\nτ/bracketleftbig\nΓρ(n,n′;ω,q)\n−∝an}b∇acketle{tΓρ(n,n′;ω,q)∝an}b∇acket∇i}htn/bracketrightbig\n= 2πδ(n−n′). (22)\nAfter obtaining the explicit form of δgZ, we can further\nsolve theδgKfrom the following relation\n(˜∂t+vF·/vector∇)δgK+i/bracketleftbig\nb·/vector σ,δgK/bracketrightbig\n+1\nτ/bracketleftbig\nδgK−∝an}b∇acketle{tδgK∝an}b∇acket∇i}htn/bracketrightbig\n= 2iφ2(r,t1)δ(t1−t2)Is−i/bracketleftbig\nφ1(r,t1)−φ1(r,t2)/bracketrightbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ\n+1\n4τ/bracketleftig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ∝an}b∇acketle{tδgZ∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ∝an}b∇acket∇i}htn−∝an}b∇acketle{t ∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ∝an}b∇acket∇i}htnδgZ∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ\n−∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦδgZ∝an}b∇acketle{t ∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ∝an}b∇acket∇i}htn+∝an}b∇acketle{t ∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦδgZ∝an}b∇acket∇i}htn∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracketrightig\n.(23)\nWe take only the zeroth and first angular harmonics\ninto account in the Keldysh component assumed spatial\nsmoothness,\n∝an}b∇acketle{tgK(t1,t2;n,r)∝an}b∇acket∇i}htΦ≈ ∝an}b∇acketle{tgK(t1,t2;n′,r)∝an}b∇acket∇i}htΦ,n′\n+2n·∝an}b∇acketle{tn′gK(t1,t2;n′,r)∝an}b∇acket∇i}htΦ,n′. (24)\nDecomposing the fluctuating term in charge and spin\ncomponents δgK=δgK\n0+δgK·/vector σ, one can easily ob-\ntain the explicit expression of the δgK\n0which is given in\nEq. (A1). The fluctuation part δgKrelated to the spin\ncomponents fulfils the following equation,\n(˜∂t+vF·/vector∇)δgK−2b×δgK+1\nτ/bracketleftbig\nδgK−∝an}b∇acketle{tδgK∝an}b∇acket∇i}htn/bracketrightbig\n=−i/bracketleftbig\nφ1(r,t1)−φ1(r,t2)/bracketrightbig/parenleftig\n∝an}b∇acketle{tgK(t1,t2;n,r)∝an}b∇acket∇i}htΦ,n\n+2n·∝an}b∇acketle{tn′gK(t1,t2;n′,r)∝an}b∇acket∇i}htΦ,n′/parenrightig\n+1\n2τ/braceleftig\n∝an}b∇acketle{tgK\n0∝an}b∇acket∇i}htΦ,n/parenleftbig\n∝an}b∇acketle{tδgZ∝an}b∇acket∇i}htn−δgZ/parenrightbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ,n+∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ,n/parenleftbig\n∝an}b∇acketle{tδgZ∝an}b∇acket∇i}htn−δgZ/parenrightbig\n∝an}b∇acketle{tgK\n0∝an}b∇acket∇i}htΦ,n/bracerightig\n.(25)\nIf denoting\nQ=\nδgK\nx\nδgK\ny\nδgK\nz\n, Lk=\ngK\nx\ngK\ny\ngK\nz\n,\nwe can write Eq. (25) in the following matrix equation,\n(˜∂t+vF·/vector∇)Q−2ζQ+1\nτ/parenleftbig\nQ−∝an}b∇acketle{tQ∝an}b∇acket∇i}htn/parenrightbig\n=−i/bracketleftbig\nφ1(r,t1)−φ1(r,t2)/bracketrightbig/parenleftbig\n∝an}b∇acketle{tLk∝an}b∇acket∇i}htΦ,n+2n·∝an}b∇acketle{tn′Lk∝an}b∇acket∇i}htΦ,n′/parenrightbig\n+1\n2τ/braceleftig\n∝an}b∇acketle{tgK\n0∝an}b∇acket∇i}htΦ,n/parenleftbig\n∝an}b∇acketle{tδgZ∝an}b∇acket∇i}htn−δgZ/parenrightbig\n∝an}b∇acketle{tLk∝an}b∇acket∇i}htΦ,n\n+∝an}b∇acketle{tLk∝an}b∇acket∇i}htΦ,n/parenleftbig\n∝an}b∇acketle{tδgZ∝an}b∇acket∇i}htn−δgZ/parenrightbig\n∝an}b∇acketle{tgK\n0∝an}b∇acket∇i}htΦ,n/bracerightig\n, (26)where the matrix ζis given by\nζ=\n0 0 −αpFcosθ\n0 0 −αpFsinθ\nαpFcosθ αpFsinθ0\n.\nThen equation (26) can be solved by utilizing the follow-\ning expression\n(−iω+ivFn·q−2ζ)Γs(n,n′;ω,q)+1\nτ/bracketleftbig\nΓs(n,n′;ω,q)\n−∝an}b∇acketle{tΓs(n,n′;ω,q)∝an}b∇acket∇i}htn/bracketrightbig\n= 2πδ(n−n′). (27)\nWe give the explicit expression of δgKin Eq. (A2) in the\nappendix A.\nSince the concrete forms of δgZandδgKhave been\nobtained, we can write down the kinetic equation satis-6\nfied by the Keldysh function through averaging the K\ncomponent of Eq. (14) over the auxiliary bosonic fields,\n(˜∂t+vF·/vector∇)∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ+i/bracketleftbig\nb·/vector σ,∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracketrightbig\n=Cel/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n+Cin/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n, (28)\nwhere the inelastic collision integral reads\nCin{∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ}(t1,t2;n,r)\n=−i∝an}b∇acketle{t/bracketleftbig\nφ1(r,t1)−φ1(r,t2)/bracketrightbig\nδgK∝an}b∇acket∇i}htΦ,(29)and the elastic collision integral is given by\nCel/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n(t1,t2;n,r) =1\nτ/bracketleftbig\n∝an}b∇acketle{t ∝an}b∇acketle{tgK(t1,t2;n,r)∝an}b∇acket∇i}htΦ∝an}b∇acket∇i}htn−∝an}b∇acketle{tgK(t1,t2;n,r)∝an}b∇acket∇i}htΦ/bracketrightbig\n+/integraldisplay\ndt3dθ1\n2π/bracketleftbig\n∝an}b∇acketle{tgK(t1,t3;n1,r)∝an}b∇acket∇i}htΦΛA(t3,t2;n1,n;r)−∝an}b∇acketle{tgK(t1,t3;n,r)∝an}b∇acket∇i}htΦΛA(t3,t2;n,n1;r)/bracketrightbig\n+/integraldisplay\ndt3dθ1\n2π/bracketleftbig\nΛR(t1,t3;n,n1;r)∝an}b∇acketle{tgK(t3,t2;n1,r)∝an}b∇acket∇i}htΦ−ΛR(t1,t3;n1,n;r)∝an}b∇acketle{tgK(t3,t2;n,r)∝an}b∇acket∇i}htΦ/bracketrightbig\n, (30)\nwhere\nΛA(t1,t2;n,n1;r) =1\n4τ/integraldisplay\ndt3∝an}b∇acketle{t/bracketleftbig\nδgZ(t1,t3;n1,r)−δgZ(t1,t3;n,r)/bracketrightbig\nδgK(t3,t2;n,r)∝an}b∇acket∇i}htΦ,\nΛR(t1,t2;n,n1;r) =1\n4τ/integraldisplay\ndt3∝an}b∇acketle{tδgK(t1,t3;n,r)/bracketleftbig\nδgZ(t3,t2;n1,r)−δgZ(t3,t2;n,r)/bracketrightbig\n∝an}b∇acket∇i}htΦ. (31)\nSubstituting the explicit forms of δgZandδgKgiven in\nEq. (21), (A1) and (A2) into Eq. (28), one get the ki-\nnetic equation which can be used to study the influence\nof electron-electron interaction on the spin dynamics of\n2DEGs with Rashba spin-orbit coupling. After some te-\ndious calculation, we obtain the explicit expressions of\nthe inelastic and elastic collision integrals, respectively,\nwhich are given in Eq. (A3) and (A6) in the appendix A.\nIV. SPIN DYNAMICS\nAfter taking average over the direction of the momen-\ntumn, one can see from Eq. (29-30) that the elastic col-\nlision integral vanishes,\n/integraldisplaydθ\n2πCel/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n(t,ǫ;n,r) = 0, (32)\nbut the average of the inelastic collision integral over the\ndirection does not vanish. This means that the elastic\ncollision integral preserves the number of electrons on a\ngivenenergyshell defined by Eq. (A5), while the inelastic\ncollision integral does not preserve it. When t1=t2,\nequation (A4) gives rise to\n/integraldisplay\ndǫCin/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n(t1,ǫ;n,r) =Cin/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n(t1,t1;n,r).One can see from Eq. (29) that the right-hand side is\nalways zero. Thus we obtain\n/integraldisplay\ndǫCin/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n(t,ǫ;n,r) = 0. (33)\nThis implies that not only the total number of electrons\nis conserved, but also the number of electrons moving\nalong a concrete direction nis conserved.\nDecomposing the Green’s function in charge and spin\ncomponents in the approximation of Eq. (24), separat-\ning the zeroth and first angular harmonics and utilizing\nEq. (32) and Eq. (33), we obtain from Eq. (28) that\nvFn·/vector∇∝an}b∇acketle{tgK(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n+i/bracketleftbig\nb·/vector σ,\n∝an}b∇acketle{tgK(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n/bracketrightbig\n=Cel/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n(t,ǫ;n,r),(34)\n∂t∝an}b∇acketle{tgK(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n+i/bracketleftbig\nb·/vector σ,2n·∝an}b∇acketle{tn′gK(t,ǫ;n′,r)∝an}b∇acket∇i}htΦ,n′/bracketrightbig\n+vFn·/vector∇/parenleftbig\n2n·∝an}b∇acketle{tn′gK(t,ǫ;n′,r)∝an}b∇acket∇i}htΦ,n′/parenrightbig\n= 0. (35)\nThere is no contribution of the inelastic collision integral\nto the spin dynamics due to the condition Eq. (33). Solv-\ning Eq. (34) and substituting ∝an}b∇acketle{tn′gK(t,ǫ;n′,r)∝an}b∇acket∇i}htΦ,n′into\nEq. (35), we can obtain the spin and charge diffusion\nequations. The diffusion equation for the charge density\nreads,\n∂tρ−CD∂2\nXρ= 0, (36)7\nwhere∂X= (∂x,∂y) andCD=v2\nFτ/2. We introduce\nthe distribution function fwhich reduces to the Fermi\ndistribution in equilibrium,\nf=f0+/vector σ·fk=1\n2(1−1\n2gK). (37)\nIn the time τ, the charge density becomes isotropic but\nthe spin relaxation process does not start yet, hence [30]\ngK\n0(ǫ) = 2(1−2f0(ǫ)),\ngK(t,ǫ;r) =−4fk(t,ǫ;r), (38)\nwhere\nf0(ǫ) =/parenleftbig\nf+(ǫ)+f−(ǫ)/parenrightbig\n/2,\nfk(t,ǫ;r) =/parenleftbig\nf+(ǫ)−f−(ǫ)/parenrightbig\ns(t,r),\nf±(ǫ) = [exp(ǫ∓∆µ/2\nkBT)+1]−1,(39)\nwheres= (sx,sy,sz) denotes the unit vector along the\nspin,f±(ǫ) represent the distribution functions projected\nalongthe directionparallelorantiparalleltothe unit vec-\ntors(all the energiesare counted from the Fermi energy)\nand ∆µ= (µ+−µ−) refers to the difference between the\nchemical potentials µ±of the electron spin subsystems.\nThediffusionequationsforthespincomponentsaregiven\nby\n∂tSx−CD∂2\nXSx−2CE∂xSz+1\nτ′sSx\n=1\nτexxSx+Fx(Sx,Sy,Sz),\n∂tSy−CD∂2\nXSy−2CE∂ySz+1\nτ′sSy\n=1\nτeyySy+Fy(Sx,Sy,Sz),\n∂tSz−CD∂2\nXSz+2CE∂xSx+2CE∂ySy+2\nτ′sSz\n=1\nτezzSz+Fz(Sx,Sy,Sz), (40)\nwhereCE=αvFpFτ,τ′\ns= 1/[2(αpF)2τ] and\nFℓ(Sx,Sy,Sz) is a quadratic form of ( Sx,Sy,Sz) lack-\ning of theS2\nℓ(ℓ=x,y,z) term. The characteristic times\nτe\nℓℓdescribe the effect of the electron-electron interaction\non the spin relaxation, and their explicit expressions aregiven by\n1\nτexx=2(αpF)2τ\nM/braceleftig/integraldisplay\ndǫ/integraldisplaydω\n2π/bracketleftbig/parenleftbig\nf+(ǫ−ω)−f−(ǫ−ω)/parenrightbig\n×Im(Rxx\n2)gK\n0(ǫ)+/parenleftbig\nf+(ǫ)−f−(ǫ)/parenrightbig\nRxx\n1gK\n0(ǫ−ω)/bracketrightbig/bracerightig\n,\n1\nτeyy=2(αpF)2τ\nM/braceleftig/integraldisplay\ndǫ/integraldisplaydω\n2π/bracketleftbig/parenleftbig\nf+(ǫ−ω)−f−(ǫ−ω)/parenrightbig\n×Im(Ryy\n2)gK\n0(ǫ)+/parenleftbig\nf+(ǫ)−f−(ǫ)/parenrightbig\nRyy\n1gK\n0(ǫ−ω)/bracketrightbig/bracerightig\n,\n1\nτezz=1\nτexx+1\nτeyy\n=2(αpF)2τ\nM/braceleftig/integraldisplay\ndǫ/integraldisplaydω\n2π/bracketleftbig/parenleftbig\nf+(ǫ−ω)−f−(ǫ−ω)/parenrightbig\n×/parenleftbig\nIm(Rxx\n2)+Im(Ryy\n2)/parenrightbig\ngK\n0(ǫ)\n+/parenleftbig\nf+(ǫ)−f−(ǫ)/parenrightbig/parenleftbig\nRxx\n1+Ryy\n1/parenrightbig\ngK\n0(ǫ−ω)/bracketrightbig/bracerightig\n,\n(41)\nwhereM=/integraldisplay\ndǫ/parenleftbig\nf+(ǫ)−f−(ǫ)/parenrightbig\n. In order to obtain the\nconcrete expressions of the characteristic times τe\nℓℓ, we\nfirstly take the energy integration in Eq. (41). Since the\nspin splitting is small, i.e.,\n|µ+−µ−|≪|µ+|,|µ−|,\nthe energy integration can be taken as follows,\n1\nM/integraldisplay∞\n−∞dǫ/parenleftbig\nf+(ǫ−ω)−f−(ǫ−ω)/parenrightbig/parenleftbig\nf+(ǫ)+f−(ǫ)/parenrightbig\n≈2/integraldisplay∞\n−∞dǫ∂f0(ǫ−ω)\n∂ǫf0(ǫ)\n/integraldisplay∞\n−∞dǫ∂f0(ǫ)\n∂ǫ\n= 1−∂\n∂ω(ωcothω\n2kBT). (42)\nAfter the energy integration, the characteristic times τe\nℓℓ\nhave the forms\n1\nτexx=1\nτeyy= 8(αpF)2τ/integraldisplay∞\n0dω\n2π/bracketleftbig∂\n∂ω(ωcothω\n2kBT)/bracketrightbig\n×/bracketleftbig\nIm(Rxx\n2)−Rxx\n1/bracketrightbig\n, (43)\nthe detail of the calculationsof the kernels Rı\n1and ImRı\n2\nare given in appendix B. Now we discuss the influence of\nthe electron-electron interaction on the spin-relaxation\ntime in the ballistic regime Tτ≫1.\nWe can obtain the characteristic time τe\nxxin the ballis-8\ntic regime utilizing the kernels Rı\n1and ImRı\n2in Eq. (B7)\n1\nτexx(Tτ≫1) = 8(αpF)2τ/integraldisplay∞\n0dω\n2π/bracketleftbig∂\n∂ω(ωcothω\n2kBT)/bracketrightbig\n×/bracketleftbig−1\n4πνv2\nF(3π\n2+tan−1ωτ−2ωτ\n1+ω2τ2)/bracketrightbig\n≈−4(αpF)2τ\nνv2\nF/integraldisplay∞\n0dω\n2π/bracketleftbig∂\n∂ω(ωcothω\n2kBT)/bracketrightbig\n=2(αpF)2τ\nπνv2\nF(2kBT−EFcothEF\n2kBT), (44)\nwheretan−1(ωτ)isreplacedby π/2forωτ≫1inthebal-\nlistic regime and EFis in the place of the upper limit of\nthe integral. In the low temperature regime kBT≪EF,\nthe second term approaches a constant independent of\nthe temperature, so the first term manifests the temper-\nature effect in the contribution of the electron-electron\ninteraction to the spin-relaxation time.\nWhen the total spin density Sis spatially homo-\ngeneous and parallel to the ℓth-axis of the coordi-\nnate frame, the contribution of Fℓ(Sx,Sy,Sz) vanishes,\nnamelyFℓ(Sx,Sy,Sz) = 0. The diffusion equations for\nspin components Sℓcan be simplified, for example,\n∂tSx=−1\nτ′sSx+1\nτexxSx=−1\nτsxxSx,(45)\nwhereτs\nxx=τ′\ns/(1−τ′\ns\nτexx). Therefore, the spin-relaxation\ntimes can be determined by τ′\nsandτe\nℓℓ, consequently,\nτs\nxx=τ′\ns\n1−τ′\ns\nτexx, τs\nyy=τ′\ns\n1−τ′\ns\nτeyy,\n(τs\nzz)−1= (τs\nxx)−1+(τs\nyy)−1. (46)\nWe can see that the total spin decays exponentially when\n0< τ′\ns/τe\nℓℓ<1. In terms of the explicit forms of\nthe characteristic times τe\nℓℓin the ballistic regime, the\nspin-relaxation times involving the effect of the electron-\nelectron interaction take the following forms\nτs\nxx=τs\nyy= 2τs\nzz=τ′\ns\n1−(T\nTF−1\n2), Tτ≫1,(47)\nwhereTF=EF/kBis the Fermi temperature. It is\nworthwhile to indicate that there exists an obvious en-\nhancement of the spin-relaxation time with increment of\nthe temperature in the ballistic regime. The increasingamplitude of the spin-relaxation time depends on the ra-\ntio of the temperature to the Fermi temperature. In con-\nclusion, an obvious enhancement of the spin-relaxation\ntime can be induced by the electron-electron interaction\nin the ballistic regime for systems under consideration.\nV. SUMMARY\nIn the above, we presented a theoretical study of the\ninfluence of electron-electron interactions on the spin dy-\nnamics for 2DEGs with Rashba spin-orbit coupling. We\nemployed the path-integral approach and the quasiclas-\nsical Green’s function to deal with the electron-electron\ninteraction. With the help of the auxiliary Bose field,\nthe electron-electron interaction was decoupled via the\nHubbard-Stratonovich transformation. Then one is able\nto derive the Elienberger equation by using the Green’s\nfunction after the transformation. Through tedious cal-\nculation, we further derived the spin and charge diffu-\nsion equations, from which the spin-relaxation time can\nbe given explicitly. We analyzed the influence of the\nelectron-electron interaction on the spin-relaxation time\nin the ballistic regime and found an obvious enhance-\nment of the spin-relaxation time with the increment of\nthe temperature T. The increasing amplitude of the\nspin-relaxation time depends on the ratio of the temper-\nature to the Fermi temperature. The electron-electron\ninteraction changes the wave vector kand hence results\nin the variation of the spin precession vector. This ex-\nhibits that the electron-electron interaction plays an im-\nportant role in the spin relaxation of electrons when\nthe D’yakonov-Perel spin relaxation mechanism domi-\nnates. It is expected to be helpful for understanding\nthe spin dynamics of 2DEGs with spin-orbit couplings\nand electron-electron interactions. Our formulation can\nalso be extend to the case of bulk inversion asymme-\ntry, namely the additional Dresselhaus term [31] with\nb=β(px,−py)+γ(pxp2\ny−pyp2\nx).\nAcknowledgments\nTheworkwassupportedbyNSFCGrantNo. 10674117\nand partially by PCSIRT Grant No. IRT0754.\nAPPENDIX A: EXPLICIT FORMS\nThe explicit form of δgK\n0is9\nδgK\n0(t1,t2;n,r) =−i/integraldisplay\ndtθ/bracketleftbig\nφ1(r1,t1−tθ)−φ1(r1,t2−tθ)/bracketrightbig/integraldisplaydθ′\n2πΓρ(tθ,n,n′;r,r1)\n×/braceleftig\n∝an}b∇acketle{tgK\n0(t1−tθ,t2−tθ;n1,r)∝an}b∇acket∇i}htΦ,n1+2n′·∝an}b∇acketle{tn1gK\n0(t1−tθ,t2−tθ;n1,r)∝an}b∇acket∇i}htΦ,n1/bracerightig\n+/integraldisplaydθ′\n2πdθ′′\n2π/integraldisplay\nd2r1dtθΓρ(tθ,n,n′;r,r1)/braceleftig\n2iφ2(r1,t1−tθ)δ(t1−t2)\n+i\nτ/bracketleftbig\n∝an}b∇acketle{tΓρ(t4−t3,n′′,n1;r2,r1)∝an}b∇acket∇i}htn1−Γρ(t4−t3,n′′,n;r2,r1)/bracketrightbig\nφ2(r2,t4)\n×/bracketleftbig\n∝an}b∇acketle{tgK\n0(t1−tθ,t3;n1,r)∝an}b∇acket∇i}htΦ,n1∝an}b∇acketle{tgK\n0(t3,t2−tθ;n1,r)∝an}b∇acket∇i}htΦ,n1\n+∝an}b∇acketle{tgK(t1−tθ,t3;n1,r)∝an}b∇acket∇i}htΦ,n1·∝an}b∇acketle{tgK(t3,t2−tθ;n1,r)∝an}b∇acket∇i}htΦ,n1/bracketrightbig/bracerightig\n. (A1)\nThe explicit expression of δgKis\nQ(t1,t2;n,r) =−i/integraldisplay\ndtθ/bracketleftbig\nφ1(r1,t1−tθ)−φ1(r1,t2−tθ)/bracketrightbig/integraldisplaydθ′\n2πΓs(tθ,n,n′;r,r1)\n×/braceleftig\n∝an}b∇acketle{tLk(t1−tθ,t2−tθ;n1,r)∝an}b∇acket∇i}htΦ,n1+2n′·∝an}b∇acketle{tn1Lk(t1−tθ,t2−tθ;n1,r)∝an}b∇acket∇i}htΦ,n1/bracerightig\n+2i\nτ/integraldisplaydθ′\n2πdθ′′\n2π/integraldisplay\nd2r1dtθΓs(tθ,n,n′;r,r1)∝an}b∇acketle{tgK\n0(t1−tθ,t3;n1,r)∝an}b∇acket∇i}htΦ,n1\n×/braceleftig/bracketleftbig\n∝an}b∇acketle{tΓρ(t4−t3,n′′,n1;r2,r1)∝an}b∇acket∇i}htn1−Γρ(t4−t3,n′′,n1;r2,r1)/bracketrightbig\n×φ2(r4,t2)∝an}b∇acketle{tLk(t3,t2−tθ;n1,r)∝an}b∇acket∇i}htΦ,n1/bracerightig\n. (A2)\nThe inelastic collision integral reads\nCin/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n(t,ǫ;n,r)\n=i\n2τ/integraldisplay\nd2r1d2r2/integraldisplaydω\n2π/bracketleftbig\nDR(ω;r,r2)−DA(ω;r2,r)/bracketrightbig\n×/bracketleftbig\n∝an}b∇acketle{tΓρ(ω;r2,r1)∝an}b∇acket∇i}ht∝an}b∇acketle{tΓρ(−ω;r,r1)∝an}b∇acket∇i}ht−∝an}b∇acketle{tΓρ(−ω;r,r1)Γρ(ω;r2,r1)∝an}b∇acket∇i}ht/bracketrightbig\n×/bracketleftbig\n∝an}b∇acketle{tgK\n0(t,ǫ−ω;n,r)∝an}b∇acket∇i}htΦ,n∝an}b∇acketle{tgK\n0(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n+∝an}b∇acketle{tgK(t,ǫ−ω;n,r)∝an}b∇acket∇i}htΦ,n·∝an}b∇acketle{tgK(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n/bracketrightbig\n−i\n2/integraldisplay\nd2r1/integraldisplaydω\n2πDK(ω;r,r1)σm/bracketleftbig\n∝an}b∇acketle{tΓs(−ω;r,r1)∝an}b∇acket∇i}ht+∝an}b∇acketle{tΓs(ω;r1,r)∝an}b∇acket∇i}ht/bracketrightbig/bracketleftbig\n∝an}b∇acketle{tLk(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n−∝an}b∇acketle{tLk(t,ǫ−ω;n,r)∝an}b∇acket∇i}htΦ,n/bracketrightbig\n+i\nτ/integraldisplay\nd2r1d2r2/integraldisplaydω\n2π/bracketleftbig\nDR(ω;r,r2)−DA(ω;r2,r)/bracketrightbig\nσm\n×/bracketleftbig\n∝an}b∇acketle{tΓs(−ω;r2,r1)∝an}b∇acket∇i}ht∝an}b∇acketle{tΓρ(−ω;r,r1)∝an}b∇acket∇i}ht−∝an}b∇acketle{tΓs(−ω;r,r1)Γρ(ω;r2,r1)∝an}b∇acket∇i}ht/bracketrightbig\n∝an}b∇acketle{tgK\n0(t,ǫ−ω;n,r)∝an}b∇acket∇i}htΦ,n∝an}b∇acketle{tLk(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n,(A3)\nwhere\nCin/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n(t1,t2;n,r)\n=/integraldisplaydǫ\n2πCin/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig/parenleftbigt1+t2\n2,ǫ;n,r/parenrightbig\neiǫ(t2−t1),(A4)\nand∝an}b∇acketle{tΓρ(s)∝an}b∇acket∇i}htmeans angular averaging defined in\nEq. (A10), the matrix σm= (σx,σy,σz) and the tem-\nporal transformation of the Green’s function has beenused due to a much faster dependence on the difference\nt1−t2than on the t1+t2\ngK(t1,t2;n,r) =/integraldisplaydǫ\n2πgK/parenleftbigt1+t2\n2,ǫ;n,r/parenrightbig\neiǫ(t2−t1),(A5)\nthe propagators of auxiliary fields have the same trans-\nformation. The elastic collision integral can be written\nas10\nCel/braceleftbig\n∝an}b∇acketle{tgK∝an}b∇acket∇i}htΦ/bracerightbig\n(t,ǫ;n,r) =−2\nτnı∝an}b∇acketle{tnıgK(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n−2\nτ/integraldisplaydω\n2πnıRı\n1(ω)∝an}b∇acketle{tgK\n0(t,ǫ−ω;n,r)∝an}b∇acket∇i}htΦ,n\n×∝an}b∇acketle{tngK(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n+i\nτ/integraldisplaydω\n2πnıRı\n2(ω)∝an}b∇acketle{tngK(t,ǫ−ω;n,r)∝an}b∇acket∇i}htΦ,n∝an}b∇acketle{tgK(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n\n−i\nτ/integraldisplaydω\n2πnı/parenleftbig\nRı\n2(ω)/parenrightbig⋆∝an}b∇acketle{tgK(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n∝an}b∇acketle{tngK(t,ǫ−ω;n,r)∝an}b∇acket∇i}htΦ,n\n+i\nτ/integraldisplaydω\n2πnı/bracketleftig\nσmRı\n3(ω)∝an}b∇acketle{tLk(t,ǫ−ω;n,r)∝an}b∇acket∇i}htΦ,n∝an}b∇acketle{tngK(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,n\n−∝an}b∇acketle{tngK(t,ǫ;n,r)∝an}b∇acket∇i}htΦ,nσm/parenleftbig\nRı\n3(ω)/parenrightbig⋆∝an}b∇acketle{tLk(t,ǫ−ω;n,r)∝an}b∇acket∇i}htΦ,n/bracketrightig\n. (A6)\nWherenı()referstothe ı()componentoftheunitvector\nn, withı,=x,yand the kernels Rı\n1(ω)−Rı\n3(ω) in\nEq. (A6) are defined by\nRı\n1(ω) = Im/integraldisplayd2q\n(2π)2DR(ω;q)/braceleftbig\n∝an}b∇acketle{tΓρ(n,ω;q)n∝an}b∇acket∇i}ht\n×∝an}b∇acketle{tnıΓρ∝an}b∇acket∇i}ht−1\n2δı,/parenleftbig\n∝an}b∇acketle{tΓρ∝an}b∇acket∇i}ht∝an}b∇acketle{tΓρ∝an}b∇acket∇i}ht−∝an}b∇acketle{tΓρΓρ∝an}b∇acket∇i}ht/parenrightbig/bracerightbig\n,(A7)\nRı\n2(ω) =/integraldisplayd2q\n(2π)2DR(ω;q)/braceleftbig\n∝an}b∇acketle{tΓρ∝an}b∇acket∇i}ht∝an}b∇acketle{tnıΓρn∝an}b∇acket∇i}ht\n−∝an}b∇acketle{tΓρnıΓρn∝an}b∇acket∇i}ht−∝an}b∇acketle{tΓρnı∝an}b∇acket∇i}ht∝an}b∇acketle{tΓρn∝an}b∇acket∇i}ht/bracerightbig\n, (A8)\nRı\n3(ω) =/integraldisplayd2q\n(2π)2DR(ω;q)/braceleftbig\n∝an}b∇acketle{tΓρn∝an}b∇acket∇i}ht∝an}b∇acketle{tnıΓs∝an}b∇acket∇i}ht\n−1\n2δı,/parenleftbig\n∝an}b∇acketle{tΓρ∝an}b∇acket∇i}ht∝an}b∇acketle{tΓs∝an}b∇acket∇i}ht−∝an}b∇acketle{tΓρΓs∝an}b∇acket∇i}ht/parenrightbig/bracerightbig\n, (A9)\nwhere we have introduced the notation\n∝an}b∇acketle{tfΓρ(s)h∝an}b∇acket∇i}ht=/integraldisplaydθdθ′\n(2π)2f(n)Γρ(s)(n,n′;ω,q)h(n′).(A10)\nAPPENDIX B: CALCULATION OF THE\nKERNELS Rı\ni\nAccording to the definition of the diffusion propagator\nΓρin Eq. (22), we can obtain\nΓρ(n,n′;ω,q) = 2πδ(n−n′)Γρ0(n;ω,q)\n+ Γρ0(n;ω,q)Γρ0(n′;ω,q)1\nτ−1\nΥ,\n(B1)\nwhere\nΓρ0(n;ω,q) =1\n−iω+ivFn·q+1\nτ\n=1\n−iω+ivFqcos(φ−φq)+1\nτ,\nΥ =/radicalbigg\n(−iω+1\nτ)2+v2\nFq2, (B2)withφqbeing the angle between the wave vector qand\nthex-axis. In terms of the explicit form of the diffusion\npropagator, one can obtain\n∝an}b∇acketle{tΓρ∝an}b∇acket∇i}ht=τ\nΥτ−1,\n∝an}b∇acketle{tΓρnx∝an}b∇acket∇i}ht=∝an}b∇acketle{tnxΓρ∝an}b∇acket∇i}ht=τ\nivFq(Υτ−1)(Υ+iω−1\nτ)cosφq,\n∝an}b∇acketle{tΓρny∝an}b∇acket∇i}ht=∝an}b∇acketle{tnyΓρ∝an}b∇acket∇i}ht=τ\nivFq(Υτ−1)(Υ+iω−1\nτ)sinφq,\n∝an}b∇acketle{tΓρΓρ∝an}b∇acket∇i}ht=−iω+1\nτ\nΥ(Υ−1\nτ)2,\n∝an}b∇acketle{tΓρnxΓρ∝an}b∇acket∇i}ht=1\nΥsin2φq\n−Υ\nv2\nFq2(Υτ−1)(1−−iω+1\nτ\nΥ)2cos2φq,\n∝an}b∇acketle{tΓρnyΓρ∝an}b∇acket∇i}ht=1\nΥcos2φq\n−Υ\nv2\nFq2(Υτ−1)(1−−iω+1\nτ\nΥ)2sin2φq,\n∝an}b∇acketle{tΓρnxΓρnx∝an}b∇acket∇i}ht=τ\nΥτ−1/parenleftbig−iω+1\nτ\nΥ2sin2φq\n−Υ−(iω+1\nτ)\nΥ2(Υτ−1)cos2φq/parenrightbig\n,\n∝an}b∇acketle{tΓρnyΓρny∝an}b∇acket∇i}ht=τ\nΥτ−1/parenleftbig−iω+1\nτ\nΥ2cos2φq\n−Υ−(iω+1\nτ)\nΥ2(Υτ−1)sin2φq/parenrightbig\n.\n(B3)\nUtilizing above formulas, we find that the kernels Rı\n1(ω)\nandRı\n2(ω) are diagonal, Rı\ni=δıRi, which can be writ-11\nten as\nR1(ω) =−Im/integraldisplay∞\n0qdq\n4πDR(ω;q)/braceleftig1\nv2\nFq2/parenleftbigΥ+iω−1\nτ\nΥ−1\nτ/parenrightbig2\n+Υ+iω−1\nτ\nΥ(Υ−1\nτ)2/bracerightig\n,\nImR2(ω) = Im/integraldisplay∞\n0qdq\n4πDR(ω;q)/braceleftig1\nv2\nFq2[Υ+iω−1\nτ]2\nΥ(Υ−1\nτ)\n+Υ+iω−1\nτ\nΥ(Υ−1\nτ)2/bracerightig\n.(B4)\nIt is not difficult to calculate the concrete forms of the\nelectron polarization operators from Eq. (18), for exam-\nple,\nΠR(ω,q) =ν[1+iω\nΥ−1\nτ]. (B5)\nSubstituting the polarization operator into Eq. (12), weobtain the propagator of the Bose fields, i.e.,\nDR(ω,q) =DR\n0\n1−DR\n0ΠR=−2πe2/q\n1+2πe2\nqν[1+iω\nΥ−1\nτ]\n≈ −1\nΠR=−1\nνΥ−1\nτ\nΥ−1\nτ−iω, (B6)\nwhere the approximation in the second line corresponds\ntotheunitarylimitassociatingwithlargerdistancesthan\nthe screening radius.\nWe obtain the concrete expressions of the kernels R1\nandR2in the ballistic regime Tτ≫1,\nR1(Tτ≫1)∝1\n8πνv2\nF(3π\n2+tan−1ωτ−2ωτ\n1+ω2τ2),\nImR2(Tτ≫1)∝ −1\n8πνv2\nF(3π\n2+tan−1ωτ−2ωτ\n1+ω2τ2).\n(B7)\n[1] S. A. Wolf, D. D. Awswchalom, R. A. Buhrman, J.\nM. Daughton, S. Von. Molnar, M. L. Roukes, A. Y.\nChtchelkanova, D. M. Tresger, Science 294, 1488 (2001),\nand references therein.\n[2] I. Zutic, J. Fabian and S. Das Sarma, Rev. Mod. Phys.\n76, 323 (2004), and references therein.\n[3] S. Datta and B. Das, Appl. Phys. Lett. 56, 665 (1990).\n[4] J. Schliemann, J. C. Egues and Daniel Loss, Phys. Rev.\nLett.90, 146801 (2003).\n[5] S. Murakami, N. Nagaosa and S. C. Zhang, Science 301,\n1348 (2003).\n[6] G. E. Pikus and A. N. Tikov, in Optical Orientation,\nModern Problems in Condensed Matter Science, Vol.8\nedited by F. Meier and B. P. Zakharchenya (North-\nHolland, Amsterdam), p109.\n[7] N. S. Averkiev and L. E. Golub, Phys. Rev. B 60, 15582\n(1999).\n[8] N. S. Averkiev, L. E. Golub and M. Willander, J. Phys.\nCondens. Matter 14, R271 (2002).\n[9] M. E. Flatt´ e, J. M. Byers and W. H. Lau, Spin Dynam-\nics in Semiconductors in Semiconductor spintronics and\nQuantum Computation, edited by D. D. Awschalom et\nal. (Springer, Berlin, 2002), chapter 4.\n[10] Y. Li and Y. Q. Li, J. Phys. Condens. Matter 19, 346231\n(2007).\n[11] M. I. D’yakonov and V. I. Perel, Zh. Eksp. Teor. Fiz, 38,\n362 (1973) [Sov. Phys. JETP. 38, 177 (1973)].\n[12] S. V. Kravchenko and M. P. Sarachik, Rep. Prog. Phys.\n67, 1 (2004).\n[13] O. Prus, Y. Yaish, M. Reznikov, U. Sivan and V. Pu-\ndalov, Phys. Rev. B 67, 205407 (2003).\n[14] A. A. Shashkin, S. Anissimova, M. R. Sakr, S. V.\nKravchenko,V.T. Dolgopolov andT. M.Klapwijk, Phys.\nRev. Lett. 96, 036403 (2006).\n[15] A. Punnoose and A. M. Finkel’stein, Science 310, 289(2005).\n[16] M. M. Glazov and E. L. Ivchenko, JETP Lett. 75, 403\n(2002).\n[17] D. Stich, J. Zhou, T. Korn, R. Schulz, D. Schuh, W.\nWegscheider, M.W.WuandC. Sch¨ uller, Phys.Rev.Lett.\n98, 176401 (2007).\n[18] W. J. H. Leyland, G. H. John, R. T. Harley, M. M.\nGlazov, E. L. Ivchenko, D. A. Ritchie, I. Farrer, A. J.\nShields and M. Henini, Phys. Rev. B 75, 165309 (2007).\n[19] M. W. Wu and C. Z. Ning, Eur. Phys. J. B 18, 373\n(2002).\n[20] M. Q. Weng and M. W. Wu, Phys. Rev. B 68, 075312\n(2003).\n[21] M. Q. Weng, M. W. Wu and L. Jiang, Phys. Rev. B 69,\n245320 (2004).\n[22] M. M. Glazov and E. L. Ivchenko, JETP 99, 1279 (2004).\n[23] A. Punnoose and A. M. Finkel’stein, Phys. Rev. Lett.\n96, 057202 (2006).\n[24] N. Nagaosa, Quantum Field Theory in Condensed Mat-\nter Physics, translated by S. Heusler (Springer, Berlin,\n1999), p95.\n[25] G. Zala, B. N. Narozhny and I. L. Aleiner, Phys. Rev. B\n64, 214204 (2001).\n[26] G. Eilenberger, Z. Phys. B 214,195 (1968).\n[27] J. Rammer and H. Smith, Rev. Mod. Phys. 58, 323\n(1986).\n[28] P. Schwab and R. Raimondi, Ann. Phys. 12, 471 (2003).\n[29] Note that the average over a variable means to ingtegrat e\nthe variable out simply, while the average over a func-\ntionalφalwayimplies anadditionalmeassureofFeymann\npath-integral, e−iS[φ], in the functional intergration.\n[30] M. I. Dyakonovand V. I. Perel, Fiz. Tverd. Tela 13, 3581\n(1971) [Sov. Phys. Solid State 13, 3023 (1972)].\n[31] G. Dresselhaus, Phys. Rev. 100, 580 (1955)." }, { "title": "1007.0853v1.Radial_Spin_Helix_in_Two_Dimensional_Electron_Systems_with_Rashba_Spin_Orbit_Coupling.pdf", "content": "arXiv:1007.0853v1 [cond-mat.mes-hall] 6 Jul 2010Radial Spin Helix in Two-Dimensional\nElectron Systems with Rashba Spin-Orbit Coupling\nYuriy V. Pershin∗\nDepartment of Physics and Astronomy and USC Nanocenter,\nUniversity of South Carolina, Columbia, SC 29208, USA\nValeriy A. Slipko\nDepartment of Physics and Technology, V. N. Karazin Kharkov National University, Kharkov, Ukraine\nWe suggest a long-lived spin polarization structure, a radi al spin helix, and study its relaxation\ndynamics. For this purpose, starting with a simple and physi cally clear consideration of spin trans-\nport, we derive a system of equations for spin polarization d ensity and find its general solution in\nthe axially symmetric case. It is demonstrated that the radi al spin helix of a certain period relaxes\nslower than homogeneous spin polarization and plain spin he lix. Importantly, the spin polarization\nat the center of the radial spin helix stays almost unchanged at short times. At longer times, when\nthe initial non-exponential relaxation region ends, the re laxation of the radial spin helix occurs with\nthe same time constant as that describing the relaxation of t he plain spin helix.\nI. INTRODUCTION\nAt the present time, there is a significant inter-\nest in the field of electron spin relaxation in semicon-\nductors stimulated by possible future applications of\nspins in electronics and computing1,2. In many two-\ndimensional (2D) electron systems the leading mecha-\nnism of spin relaxation is the D’yakonov-Perel’ spin re-\nlaxation mechanism3,4. Within this mechanism, electron\nspins feel an effective momentum-dependent magnetic\nfield randomized by electron scattering events resulting\nin relaxation of electron spin polarization. A number\nof theoretical and experimental studies on peculiarities\nof D’yakonov-Perel’ spin relaxation were reported in the\nlast decade5–16.\nIt wasshownin Ref. 9that the spin relaxationtime for\n2D electrons depends not only on material parameters\n(e.g., strength of spin-orbit interaction, electron mean\nfree path, etc.) but also on the initial spin polarization\nprofile. In particular, it was demonstrated that a plain\nspin helix in a 2D electron system with Rashba spin-\norbit interaction has a longer spin relaxation time than a\nhomogeneousspinpolarization9. Alaterstudy11revealed\nthat in a system with both Rashba17and Dresselhaus18\ninteractions such an increase in spin relaxation time can\nbe even more dramatic. This effect was also observed\nexperimentally13,19.\nIn this paper, we consider spin relaxation of a radial\nspin helix in a 2D electron system with Rashba spin-orbit\n(SO) interaction. This structure is interesting because it\nprovides, to the best of our knowledge, the longest spin\nrelaxation time of 2D electrons subjected to Rashba SO\ninteraction. Such a property is related to spin polariza-\ntion in the vicinity ofthe special point of radialspin helix\nr= 0. Physically, at short times, when the period of ra-\ndial spin helix is equal to the period of spin precession,\nan electron diffusing from any direction to a point in the\nvicinity of r= 0 has the same direction of spin polariza-\ntion as the initial spin polarization at this point. Thus,D’yakonov-Perel’ spin relaxation becomes inefficient at\nshort times for spin polarization in the vicinity of r= 0.\nIn the radial spin helix, the initial distribution of spin\npolarization has a cylindrical (axial) symmetry and is\ngiven by\nSr(r,t= 0) =−S0sin(kr), (1)\nSϕ(r,t= 0) = 0 , (2)\nSz(r,t= 0) =S0cos(kr), (3)\nwherekisthe wavevector, and S0istheinitial amplitude\nof spin polarization. In Fig. 1(a) we show schematically\nspin polarizationdistribution in the radialdirection. The\noverall distribution of z-component of the spin polariza-\nz (a) z\nr(a)\nFIG. 1. (Color online) (a) Schematic of initial spin polariz a-\ntion distribution in a radial spin helix. (b) Initial distri bution\nofz-component of spin polarization ( Sz(r,0)/S0) in the radial\nspinhelix offinite radius usedin ourMonte Carlo simulation s.2\ntion in radial spin helix of finite radius is shown in Fig.\n1(b). Moreover, in this paper we will often refer to a\nplain spin helix suggested in Ref. 9. The initial distri-\nbution of spin polarization components in the plain spin\nhelix is\nSx(x,t= 0) =−S0sin(kx), (4)\nSy(x,t= 0) = 0 , (5)\nSz(x,t= 0) =S0cos(kx). (6)\nThis paper is organized as follows. In Sec. II, a drift-\ndiffusion equation approach is used to study relaxation\ndynamics of the radial spin helix. We obtain general ex-\npressions for spin polarization as a function of time and\nstudy its short and long time behavior. Our analytical\nstudies are supported by Monte Carlo simulations pre-\nsented in Sec. III. Our main results and conclusions\nare summarized in Sec. IV. Moreover, in several Ap-\npendices following the main text, we provide additional\ncalculation details. Specifically, electron spin rotations\ninduced by Rashba SO interaction are considered in the\nAppendix A, spin drift-diffusion equations are derived\nin the Appendix B, a general analytical solution of spin\ndrift-diffusion equation in the axially symmetric case is\npresentedin the Appendix C, andrelaxationofplainspin\nhelix is discussed in the Appendix D.\nII. DRIFT-DIFFUSION DESCRIPTION OF\nRADIAL SPIN HELIX\nIn this section we consider the relaxation dynamics of\nthe radial spin helix analytically. The initial spin polar-\nization in the radial spin helix is of cylindrical symmetry\nand described by Eqs. (1-3). Intuitively, a special point\nin the radial spin helix is r= 0, because the electrons\nmotion through this point along straight trajectories in\nalldirections should not lead to spin relaxation at short\ntimes for a specific value of the wave vector k. Corre-\nspondingly, the spin lifetime of electrons located in a re-\ngion within r= 0 should be longer than the spin lifetime\nof homogeneous spin polarization and of plain spin helix.\nThis effect is in the focus of our investigation.\nLet us consider a two-dimensional electrons confined\nin a quantum well or heterostructure with Rashba-type\nspin-orbit interaction17. The standard Hamiltonian with\nthe Rashba term is given by\nˆH=ˆ p2\n2m+α(ˆσ׈ p)·z, (7)\nwhereˆ p= (ˆpx,ˆpy) is the 2D electron momentum oper-\nator,mis the effective electron’s mass, ˆσis the Pauli-\nmatrix vector, αis the spin-orbit coupling constant and\nzis a unit vectorperpendicular to the confinement plane.\nIt is not difficult to show that in the case of Hamil-\ntonian (7) the quantum mechanical evolution of a spin\nof an electron with a momentum pcan be reduced to\na spin rotation with the angular velocity Ω = 2 αp//planckover2pi1about the axis determined by the unit vector n=p×z/p\n(see Appendix A). In this way, the spin-orbit coupling\nconstant αenters into equations through the parameter\nη= 2αm/planckover2pi1−1, which gives the spin precession angle per\nunit length.\nBesides this evolution, 2D electrons experience dif-\nferent bulk scattering events such as, for example, due\nto phonons or impurities. These scatterings random-\nize the electron trajectories. Correspondingly, the di-\nrection of spin rotation becomes fluctuating what causes\naverage spin relaxation (dephasing). This is the famous\nD’yakonov-Perel’spinrelaxationmechanism.3,4Thetime\nscale of the bulk scattering events can then be character-\nized by a single rate parameter, the momentum relax-\nation time τ. It is connected to the mean free path by\nℓ=vτ, where v=p/mis the mean electron velocity.\nTo take into account these scatterings we use a model\nof diffusive spin transport, which in the limit of small\nkℓ≪1, yields the spin drift-diffusion equations (B3-B5)\n(Appendix B provides derivation details).\nLet us consider dynamics of a radial spin helix relax-\nation. We assume that such a structure is created at the\ninitial moment of time with the spin polarization com-\nponents given by Eqs. (1-3). The exact solution of the\nradial spin drift-diffusion equations (C1-C2) with initial\nconditions (1-3) and constants γandCfrom Eqs. (B6)\ncan be written as (see Appendix C for more details)\nSr(r,t)\nS0=−d\ndk/integraldisplayk\n0dsJ1(sr)√\nk2−s2/bracketleftbigg\nkcosh/parenleftbigg/radicalbig\nη2+16s2ηDt\n2/parenrightbigg\n+/parenleftbig\nkη+4s2/parenrightbigsinh/parenleftBig/radicalbig\nη2+16s2ηDt\n2/parenrightBig\n/radicalbig\nη2+16s2\ne−(s2+3η2/2)Dt,(8)\nSz(r,t)\nS0=d\ndk/integraldisplayk\n0dssJ0(sr)√\nk2−s2/bracketleftbigg\ncosh/parenleftbigg/radicalbig\nη2+16s2ηDt\n2/parenrightbigg\n+(4k��η)sinh/parenleftBig/radicalbig\nη2+16s2ηDt\n2/parenrightBig\n/radicalbig\nη2+16s2\ne−(s2+3η2/2)Dt,(9)\nwhereJ1(r) andJ0(r) are the Bessel functions ofthe first\nand zeroth order correspondingly and Dis the diffusion\nconstant.\nEqs. (8-9) define completely the radial spin helix at\nany point rand at any moment of time t. The time\ndependence of spin polarization at the center of helix is\nof particular interest because the spin relaxation at this\npoint is the slowest. The radial component of spin po-\nlarization Srin the vicinity of r= 0 is close to zero (it\nfollows from symmetry considerations or directly from\nEq. (8)). Therefore, below, we derive asymptotic ex-\npressions at short and long times for Szonly. At short\ntimesDη2t≪1, an expansion of the RHS of Eq. (9) in3\ntand its integration over satr= 0 results in\nSz(0,t)\nS0= 1−2(k−η)2Dt+2\n3(k−η)(2k3−6k2η\n+6kη2−3η3)D2t2−/parenleftbigg8\n15k6−16\n5k5η+8k4η2−104\n9k3η3\n+32\n3k2η4−14\n3kη5+4\n3η6/parenrightbigg\nD3t3+O(t4).(10)\nIt follows from Eq. (10) that when the radial spin helix\nperiod is equal to spin precession length (this happens\nwhenk=η) the decay of Szat the center of helix starts\nwith a cubic term in t\nSz(0,t)\nS0= 1−10\n9(Dη2t)3+O(t4),fork=η,Dη2t≪1.\n(11)\nThis means that the spin relaxation at the center of the\nradial spin helix at short times is significantly suppressed\n(for this special wave number, k=η) and characterized\nby a rather long initial interval of non-exponential be-\nhaviour.\nThe asymptotic behaviour of SrandSzat long times,\nDη2t≫1, can be determined by taking into account the\ndominant contribution to the integrals in Eqs. (8) and\n(9). This contribution comes from the vicinity of a point\ns∈[0,k]correspondingtothemaximumofthe −λ−(s)in\nthe integration interval (see Eq. (C10) and Fig. 7). We\nshould consider three cases. In the first case, when 0 <\nk < km=√\n15η/4, the main contribution to the integrals\nin Eqs. (8) and (9) comesfrom the vicinity ofpoint s=k\nat the right end of the integration interval. In the second\ncase, when k=km, we should keep in mind that k=\nkmis a stationary point of λ−(s) entering the exponent.\nTherefore, in this case, the asymptotic behaviour differs\nfrom the case k < kmby a pre-exponential factor. In the\nthird case, when k > k m, the main contribution to the\nintegrals in Eqs. (8) and (9) arises from the vicinity of\nthe inner stationary point s=kmofλ−(s).\nThe asymptotic behaviour of Laplace-type integrals is\nobtained using standard for this purpose technics. From\nEq. (9) we get\nSz(r,t)\nS0=/parenleftBigg\n1+4k−η/radicalbig\nη2+16k2/parenrightBigg\nJ0(kr)/radicalbigg\n−πk\n8dλ−(k)\ndkt\n×e−λ−(k)t,for 0< k <√\n15\n4η,−kdλ−(k)\ndkt≫1,(12)\nSz(r,t)\nS0=√\n15(√\n15+3)\n32√\n2Γ/parenleftbigg3\n4/parenrightbigg\nJ0(kmr)(Dη2t)1\n4\n×e−λ−(km)t,fork=km=√\n15\n4η, Dη2t≫1,(13)\nSz(r,t)\nS0=−3\n32(4k+5η)η\n(k2−k2m)3/2J0(kmr)/radicalbiggπ\nDte−λ−(km)t,\nfork >√\n15\n4η,(k−km)2Dt≫1, Dη2t≫1.(14)0 1 2 3 4 5 6 7 8 9 0.0 0.2 0.4 0.6 0.8 1.0 \n k =0.5 η\n k =km\n k =1.6 η\n Sz/S 0\nDη2t12\n3\nFIG. 2. (Color online) Time-dependence of Sz(0,t) in the\nradial spin helix at several values of k.\nFrom Eq. (12) we see that if the wave vector ksatis-\nfies the inequality 0 < k < k m, then, at long times, the\nspin polarization decay is mainly exponential and char-\nacterized by a relaxation time τ(k) = (λ−(k))−1. This\nrelaxation time increases with k. At the same time (ac-\ncordingly to Eq. (10)), the spin polarization decay at\nshort times also decreases with k. Therefore, the spin\nlife time of the radial spin helix increases with increase\nofk∈(0,km).\nAtk=km, as it follows from Eqs. (13-14),\nthe relaxation time reaches its maximum value τm=\n(λ−(km))−1= (7Dη2/16)−1. Moreover, since km≈\n0.97ηis very close to η, the conditions for short time\nsuppression of spin relaxation are almost optimal at this\nvalue ofk. Therefore, when k=km, the spin relaxation\nis significantly suppressed at both short and long times.\nWhenk > km, the asymptotic relaxation time is the\nsame as when k=km(see Eqs. (13,14)). However, even\nwhenkis close to km, the ratio of the absolute value of\nRHSofEq. (14)tothoseofEq. (13)issmall(oftheorder\nof [(k−km)2Dt]−3/4). In addition, in the asymptotic\nformula (14), the pre-exponential factor changes its sign.\nTherefore, the spin polarization Szmust turn to zero at\nsome moment of time, before it reaches the asymptotic\nbehaviour given by Eq. (14). Eq. (10) at k > η≈km\nalso predicts a relaxation increase with kwhenk > km.\nThus we conclude that the longest spin relaxationtime\nfor the spin polarization at the center of the radial spin\nhelix occurrs at the wave vector\nk=km=√\n15η/4 (15)\nand is given by\nτm= (7Dη2/16)−1. (16)\nIn addition, at this value of k, the dynamics of spin po-\nlarization in the vicinity of r= 0 is non-exponential at\nshort times, when the spin polarization remains almost\nconstant. These are the main results of our calculations.4\n(a)0 2 4 6 8 10 12 -0.01 0.00 0.01 0.02 \n k= 0.6 η\n k=k m\n k= 1.4 η\n Sz/S 0\nηrDη2t= 10 \n(b)0 2 4 6 8 10 12 -0.01 0.00 0.01 k= 0.6 η\n k=k m\n k= 1.4 η\n Sr/S 0\nηrDη2t=10 \nFIG. 3. Spatial dependence of Sz(a) andSr(b) in the radial spin helix at the indicated moment of time.\nThe relaxation of initially homogeneous spin polariza-\ntion (when k= 0) can be obtained from Eq. (9) in the\nlimitk→0. In this limiting case the factor before the\nexponential function e−λ−(s)tin Eq. (9) turns to zero in\nk= 0 limit. Therefore, at k= 0, the spin relaxation is\ndetermined by the exponential function e−λ+(0)t, where,\naccordingly to Eq. (C10), λ+(0) = 2Dη2.\nIn fact, the exact time dependence of Szatk= 0\ncoincideswith its asymptoticbehaviour. It can be clearly\nseen from both Eqs. (9) and (B5) that\nSz(t) =S0e−λ+(0)t=S0e−2Dη2t,fork= 0.(17)\nAccordingly to Eq. (17), the initially homogeneous spin\npolarization, directed along z-axis, decays exponentially\nwith a time constant τ(h)= (2Dη2)−1. We also note that\nthe applicability limits of the long times asymptotic ex-\npressions listed in Eqs. (12-14) do not allow calculations\nofSzatk= 0 ork=kmas a limiting case of Eq. (12).\nIn Fig. 2 we show the time dependence of the spin\npolarization component Sz(r= 0,t) calculated at several\nwave vectors ( k= 0.5η(solid line 1), k=km=√\n15η/4\n(solid line 2) and k= 1.6η(solid line 3)) using Eq. (9).\nThe dashed lines 1, 2 and 3 represent asymptotic behav-\nior at short times (given by Eq. (10)), and dotted lines\n1, 2, 3 show asymptotic behavior at long times (given by\nEqs. (12-14)) for the same values of the wave vectors k.\nThese curves reveal main features discussed above.\nThe spatial dependences of z- andr-components of\nspin polarization, calculated from Eqs. (9) and (8), are\npresented in Fig. 3 for a particular moment of time\nDη2t= 10 and wave vectors k= 0.6η(dashed lines),\nk=km(solid lines), and k= 1.4η(dotted lines). These\nplots clearly demonstrate that the maxima of SzandSr\nare reached at the wave vector k=km.\nFig. 4 depicts time dependence of Sz(r= 0,t) for ho-\nmogeneous spin polarization (calculated from Eq. (17)),\nplain spin helix (calculated from Eq. (D4)) and radial0 2 4 6 8 10 0.0 0.2 0.4 0.6 0.8 1.0 homogeneous polarization \n plane spin helix, k=k m\n radial spin helix, k=k m\n Sz/S 0\nDη2tr = x = 0, \nkm=0.97 η\nFIG. 4. (Color online) Time dependence of Sz(r= 0,t) for\nthree different initial spin configurations: homogeneous po -\nlarization, plain spin helix and radial spin helix. The radi al\nspin helix is characterized by the largest magnitude of spin\npolarization at any moment of time.\nspin helix. This plot demonstrates that the spin polar-\nization at r= 0 in the radial spin helix lives longer that\nthose in the case of homogeneous spin polarization and\nplain spin helix. It is interesting and important that at\nshort times this curve stays almost flat as expected. At\nlonger times, when the initial non-exponential relaxation\nregion ends, the relaxation of spin polarization in the\nradial spin helix occurs with the same time constant as\nthose of the plain spin helix (it can be shown analytically\nfrom Eqs. (D3-D4) that the longest spin relaxation time\nof the plain spin helix is given by Eq. (16) and occurs\natk=kmgiven by Eq. (15)). Therefore, in both cases,\nthe increase of the exponential relaxation time relative\nto the homogeneous spin polarization is equal to 32 /7.5\n0 50 100 150 200 -1.0 -0.5 0.0 0.5 1.0 \nSz\nSrSpin polarization (arb. units) S\nr (in units of m.f.p.) (b) \nFIG. 5. (Color online) Distributions of zcomponent (a)\nand all components (b) of spin polarization ( Sϕ= 0,S=/parenleftbig\nS2\nr+S2\nz/parenrightbig1/2) att= 100τin the radial spin helix. The red\n(dark) area in the center of radial spin helix in (a) and the\nmaximum of SandSzin the vicinity of r= 0 in (b) demon-\nstrate a longer spin lifetime of electrons located in this re gion.\nThis plot was obtained at ηℓ= 0.1 and the radial spin helix\nperioda= 64.77ℓ(this value of acorresponds to k=km).\nM.f.p. (mean free path) stands for ℓ.\nIII. MONTE CARLO SIMULATIONS\nIn order to obtain an additional insight on spin relax-\nation of the radial spin helix, we perform Monte Carlo\nsimulations employing an approach described in Refs. 5\nand 20. This Monte Carlo simulation method uses a\nsemiclassical description of electron space motion and\nquantum-mechanical description of spin dynamics (the\nlaterisbasedontheHamiltonian (7)). All specific details\nof the Monte Carlo simulations program can be found in\nthe references cited above and will not be repeated here.\nTo some extent, Monte Carlo simulations program nu-\nmerically solves Eqs. (B3-B5) taking into account rela-\ntions (A4) and (A5).\nAll numerical results related to the radial spin helix\nwere obtained using an ensemble of 108electrons initially0 200 400 600 800 1000 0.01 0.1 1\n Sz(0,t) /Sz(0,0) \nTime ( τ) Radial spin helix \n Plain spin helix \n Homogeneous polarization \nFIG. 6. (Color online) Spin polarization as a function of tim e\nfor different initial spin polarization configurations. Thi s plot\nwas obtained using the parameters values ηℓ= 0.1 anda=\n64.77ℓ(the period of helices). The straight lines are fitting\ncurves selected as exp( −(t−t0)/t1). The parameters of the\nfitting curves are t0= 0,t1= 103τfor homogeneous spin\npolarization, t0= 0τ,t1= 435τfor plain spin helix, and\nt0= 165τ,t1= 430τfor radial spin helix.\nhomogeneouslydistributedwithin acircleofasufficiently\nlarge radius R= 200ℓto insure that the influence of\nboundary effects on spin polarization in a region in the\nvicinity of r= 0 is negligible. The initial configuration\nofz-component of spin polarization of these electrons is\npresented in Fig. 1(b). Our Monte Carlo simulations\nresults are in an excellent agreement with the theory of\nradial spin helix relaxation presented above.\nFig. 5(a) demonstrates the spatial dependence of Szin\ntheradialspin helixat t= 100τandFig. 5(b)depicts the\nradialdependence of the spin polarizationcomponents at\nthe same moment of time. In particular, it can be clearly\nseen that the decay of Szin the vicinity of r= 0 is\nslower than in other regions. The total spin polarization\nSshows oscillations that are related to a well-known fea-\nture of D’yakonov-Perel’ relaxation: the spin relaxation\ntime of the perpendicular to plane spin polarization is\nshorter than the spin relaxation time of the in-plane spin\npolarization. Similar oscillation of spin polarization am-\nplitude were previously found in the relaxation dynamics\nof the plain spin helix.9\nIn addition to the radial spin helix relaxation, we sim-\nulated the relaxation of homogeneous spin polarization\nand relaxation of plain spin helix. Monte Carlo simu-\nlations reveal that at long times the time dependence of\nspin polarization in all spin configurations (homogeneous\nspin polarization, plain spin helix and radial spin helix)\nexhibits an exponential decay.\nIn Fig. 6 we compare the time dependencies of spin\npolarization ( S(0,t)/S0) for three different initial polar-\nization configurations: the homogeneous spin polariza-\ntion (initial spin polarization is selected as Sx=Sy= 0,\nSz=S0), plain spin helix and radial spin helix. The6\nselected period of radial and plane spin helices ( k=km)\ncorresponds to the longest spin lifetime of these struc-\ntures at a fixed ηℓ= 0.1. It follows from Fig. 6 that\nthe spin polarization in the radial spin helix is the most\nrobust against the relaxation. We emphasize that the\nlong-time behavior of all curves can be perfectly fitted\nby an exponential law given in the caption of Fig. 6.\nThe numerically obtained increase in the spin lifetime of\nthe radial spin helix 435 /103∼4.2 is very close to the\ntheoretically predicted value 32 /7∼4.6. We also note\nthat a slightly longer spin relaxation time increase ( ∼6)\nreported in Ref. 9 for the plain spin helix can be related\nto a large value of ηℓ= 0.3 that is beyond the linear spin\ndrift-diffusion theory.\nIV. CONCLUSION\nThe dynamics of spin relaxation of radial spin helix\nwas investigated using spin drift-diffusion equations and\nMonte Carlo simulations. Starting with a clear model\nof diffusive spin transport, we derived spin drift-diffusion\nequations for electron spin polarization in 2D semicon-\nductor structures with Rashba spin-orbit coupling. The\ngeneral solution of these equations for the axial symmet-\nric case was found. Based on this solution, we studied\nthe evolution of the suggested long-lived spin structure -\nthe radial spin helix. It was shown that the relaxation of\nspin polarization in the vicinity of r= 0 in this structure\ndemonstrates an unusual long initial non-exponential re-\nlaxation behavior followed by an exponential decay. The\noptimal value of the radial spin helix wave vector was\nfound andcorrespondingexponentialrelaxationtime was\ncalculated. Qualitatively, the initial non-exponential de-\ncayfeaturecanbeexplainedbyexistenceofaninfiniteset\nof dephasing-free trajectories propagating through the\npointr= 0.\nIn ordertoadditionallycheckouranalyticalresults, we\nalso performed Monte Carlo simulations of the dynamics\nofradialspinhelixrelaxationusingthesameMonteCarlo\nsimulation technique as those described in Refs. 5 and\n20. Fig. 5 shows a representative result of our simulation\nin which it is clearly demonstrated that the polarization\ndecay at r= 0 is slowest. Our analytical and Monte\nCarlo simulations results are in a perfect agreement.\nTo conclude, the radial spin helix is a new structure\nexhibiting an unusual spin relaxation dynamics and rel-\natively long lifetime. Its property of slow relaxation dy-\nnamicsatshorttimesisveryinteresting. Experimentally,\nthe radialspin helix canbe createdbyspin injection from\na point electrode located at the center of a second ring-\nshape electrode or possibly by a modified spin gratings\ntechnique21.Appendix A: Rotation of a single spin subjected to\nRashba SO interaction\nLet us consider an electron with a momentum p=\n(px,py). The Hamiltonian (7) can be rewritten as\nˆH=p2\n2m+αpn·ˆσ, (A1)\nwheren=p×z/pis the unit vector.\nSince eigenvalues spin projection operator\nˆ s·n=ˆσ·n/2 on any direction are equal to ±1/2,\nthe energy levels of (A1) are given by\nE±=p2\n2m±αp, (A2)\nand it is clear that corresponding spin eigenfunctions are\nthe states with spin directed along and opposite to n.\nThe evolution operator acting only on spin variables of\nthe electron with momentum pis equal to\nˆU(t) =e−iˆHt//planckover2pi1= exp/braceleftbigg−ip2t\n2m/planckover2pi1/bracerightbigg\nexp/braceleftbigg\n−iΩtn·ˆσ\n2/bracerightbigg\n,\n(A3)\nwhere Ω = 2 αp//planckover2pi1.\nWe note that the evolution operator (A3) coincides\nwith the operator of finite rotation by an angle Ω t=\n2αpt//planckover2pi1aboutn-axis up to the phase factor e−ip2t/(2m/planckover2pi1).\nSince electron spin stransforms under rotations as the\nordinary vector, we obtain\ns(t) = cos(Ω t)s(0)+sin(Ω t)n×s(0)+2sin2(Ωt/2)n·s(0)n,\n(A4)\nwhere\nΩ = 2αp//planckover2pi1,n=p×z/p. (A5)\nAppendix B: Derivation of spin drift-diffusion\nequations\nLet usconsidera two-dimensionalnon-degenerateelec-\ntron gas and use a semiclassical approach to model\nthe electron space motion and quantum-mechanical ap-\nproachbased on the Hamiltonian (7) to describe the elec-\ntron spin dynamics. Moreover, we assume the electri-\ncal neutrality and absence of an external electromagnetic\nfield. Within our approach, 2D electrons are character-\nized by the momentum relaxation time τand the mean\nfree path ℓ, so that the average velocity of electrons is\nv=ℓ/τ. From elementary gas-kinetic considerations22\nwe can write an equation for the change of electron spin\npolarization ∆ S(x,y,t) in a region of dimensions 2 ℓ×2ℓ\nwith the center at ( x,y) during the time interval τ:\n(2ℓ)2∆S(x,y,t) =1\n4vτ(2ℓ){S′(x−2ℓ,y,t)\n+S′(x+2ℓ,y,t)+S′(x,y−2ℓ,t)+S′(x,y+2ℓ,t)(B1)\n−4S(x,y,t)}.7\nIn the right hand side of Eq. (B1), the first four terms\nare the spin polarization fluxes into the region from four\nsides with length 2 ℓ, and the last term is the flux out\nof this region. The prime symbols in Eq. (B1) denote\na change of spin polarization because of SO interaction-\ninduced spin precession by the angle 2Ω τ= 4αmℓ//planckover2pi1=\n2ηℓaccordingly to Eqs. (A4,A5). For example,\nS′(x−2ℓ,y,t) = cos(2 ηℓ)S(x−2ℓ,y,t)−sin(2ηℓ)y\n×S(x−2ℓ,y,t)+2sin2(ηℓ)y·S(x−2ℓ,y,t)y,(B2)\nwhereyis the unit vector along y−axis.\nIn order to obtain drift-diffusion equations for spin\npolarization, we substitute expressions for S′into Eq.\n(B1), andexpandtrigonometricalfunctionsuptoquadric\nterms with respect to small 2 ηℓandS′terms up to\nquadric terms with respect to 2 ℓ. The resulting system\nof drift-diffusion equations for spin polarization have a\nform\n∂Sx\n∂t=D∆Sx+C∂Sz\n∂x−2γSx,(B3)\n∂Sy\n∂t=D∆Sy+C∂Sz\n∂y−2γSy,(B4)\n∂Sz\n∂t=D∆Sz−C/parenleftbigg∂Sx\n∂x+∂Sy\n∂y/parenrightbigg\n−4γSz,(B5)\nwhere\nC= 2ηD, γ=1\n2η2D, (B6)\nand\nD=ℓ2\n2τ. (B7)\nHereDis the coefficient of diffusion, Cdescribes spin\nrotations, and γis the coefficient describing spin relax-\nation.\nIt is interesting to note that the same drift-diffusion\nequations (B3-B6) can be obtained for the model of 2D\nlocalized electrons on a lattice23in the hopping regime.\nHowever, in this case, the diffusion coefficient is equal to\nD=ℓ2/(4τ), where τis the characteristic hopping time\nandℓis the distance between lattice sites.\nAppendix C: Analytical solution of drift-diffusion\nequations in the axially symmetric case\nIn the axially symmetric case (assuming that Sr=\nSr(r,t),Sz=Sz(r,t) andSϕ= 0) Eqs. (B3-B5) can be\nwritten as\n∂Sr\n∂t=D/bracketleftbigg∂\nr∂r/parenleftbigg\nr∂Sr\n∂r/parenrightbigg\n−Sr\nr2/bracketrightbigg\n+C∂Sz\n∂r−2γSr,(C1)\n∂Sz\n∂t=D∂\nr∂r/parenleftbigg\nr∂Sz\n∂r/parenrightbigg\n−C∂(rSr)\nr∂r−4γSz.(C2)Let us find a general solution of Eqs. (C1-C2) for the\ncase of an infinite plane. We search a specific solution of\nthe above Eqs. in the form\nSr(r,t) =A(s,t)J1(sr), (C3)\nSz(r,t) =B(s,t)J0(sr), (C4)\nwhereJ1(r) andJ0(r) are the Bessel functions of the\nfirst and zeroth order correspondingly. Substituting ex-\npressions (C3-C4) into Eqs. (C1-C2) we obtain a system\nof ordinary differential equations for unknown functions\nA(s,t) andB(s,t) of positive parameter sand time t\ndA(s,t)\ndt=−(Ds2+2γ)A(s,t)−CsB(s,t),(C5)\ndB(s,t)\ndt=−CsA(s,t)−(Ds2+4γ)B(s,t).(C6)\nThe general solution of this system can be presented as\nA(s,t) =C+(s)sCe−λ+(s)t+C−(s)sCe−λ−(s)t,(C7)\nB(s,t) =C+(s)/parenleftBig\nγ+/radicalbig\nγ2+C2s2/parenrightBig\ne−λ+(s)t(C8)\n+C−(s)/parenleftBig\nγ−/radicalbig\nγ2+C2s2/parenrightBig\ne−λ−(s)t,\nwhere we denote\nλ±(s) =Ds2+3γ±/radicalbig\nγ2+C2s2,(C9)\nandC±(s) are arbitrary functions of positive parameter\ns. Using Eqs. (B6), Eq. (C9) can be rewritten in a more\nsimple form\nλ±(s) =1\n2D(2s2+3η2±η/radicalbig\nη2+16s2).(C10)\nThe special solutions (C3-C4) (with A(s,t) andB(s,t)\ngiven by Eqs. (C7-C9)) of the radial drift-diffusion equa-\ntions (C1-C2) have the following simple meaning. Ac-\ncordingly to Eqs. (C3-C4), the spatial dependencies of\nthe radial and z-components of spin polarization are pro-\nportional to the first and zeroth order Bessel function\nat any moment of time. The parameter sis similar\nto the wave vector kfor the plane case. The ampli-\ntudesA(s,t) andB(s,t) determine the time dependence\nof the radial and z-components of spin polarization. If\nC+(s)/ne}ationslash= 0,C−(s) = 0 ( C+(s) = 0,C−(s)/ne}ationslash= 0), then\nthese amplitudes are exponential functions of time with\nthe inverse relaxation time λ+(s) (λ−(s)) as it can be\nseen from Eqs. (C7-C8).\nWhereas the inverse relaxation time λ+(s) takes its\nminimum value at s= 0 and monotonically increases\nwith parameter s, the inverse relaxation time λ−(s) has\na minimum at\nsm=/radicalbigg\nC2\n4D2−γ2\nC2. (C11)\nAt this value of s,λ−(s) is equal to\nλm= 3γ−C2\n4D−γ2D\nC2. (C12)8\nUsing relations (B6), we find that the minimum value\nofλ−(s) is equal to\nλm=7\n16Dη2, (C13)\nand it occurs at\nsm=√\n15\n4η. (C14)\nFig. 7 shows λ±(s) given by Eq. (C10) as a function of\ns.\nIn order to obtain the general solution of Eqs. (C1-\nC2), we should integrate the special solutions (C3-C4)\nover positive parameter staking into account relations\n(C7-C9). Two arbitrary functions C±(s) entering Eqs.\n(C7-C9) can be found from specified initial conditions\nfor the radial and z-components of polarization in the\nform of its Fourier-Bessel transforms\n˜Sr(s) =/integraldisplay+∞\n0drrSr(r,0)J1(sr),(C15)\n˜Sz(s) =/integraldisplay+∞\n0drrSz(r,0)J0(sr).(C16)\nAs a result of algebraic transformations, we obtain the\nsolution of the initial value problem (C1-C2) in the case\nof the infinite plane\nSr(r,t) =/integraldisplay+∞\n0dssGrr(r,t,s)˜Sr(s)\n+/integraldisplay+∞\n0dssGrz(r,t,s)˜Sz(s),(C17)\nSz(r,t) =/integraldisplay+∞\n0dssGzr(r,t,s)˜Sr(s)\n+/integraldisplay+∞\n0dssGzz(r,t,s)˜Sz(s),(C18)\n0.0 0.5 1.0 1.5 2.0 2.5 3.002468\nλ(s)/Dη2λ+(s)\nλ−(s)\ns/η\nFIG. 7. (Color online) s-dependence of inverse relaxation\ntimesλ±(s). The minimum of λ−(s) corresponds to the wave\nvector giving the longest relaxation time.wheretheGreenfunctions Grr,Gzr,Grz,Gzzaredefined\nas follows\nGrr(r,t,s) =J1(rs)/bracketleftBig\ncosh/parenleftBig\nt/radicalbig\nγ2+C2s2/parenrightBig\n+γsinh/parenleftBig\nt/radicalbig\nγ2+C2s2/parenrightBig\n/radicalbig\nγ2+C2s2\ne−(Ds2+3γ)t,(C19)\nGrz(r,t,s) =−CJ1(rs)sinh/parenleftBig\nt/radicalbig\nγ2+C2s2/parenrightBig\n/radicalbig\nγ2+C2s2\n×se−(Ds2+3γ)t,(C20)\nGzr(r,t,s) =−CJ0(rs)sinh/parenleftBig\nt/radicalbig\nγ2+C2s2/parenrightBig\n/radicalbig\nγ2+C2s2\n×se−(Ds2+3γ)t,(C21)\nGzz(r,t,s) =J0(rs)/bracketleftBig\ncosh/parenleftBig\nt/radicalbig\nγ2+C2s2/parenrightBig\n−γsinh/parenleftBig\nt/radicalbig\nγ2+C2s2/parenrightBig\n/radicalbig\nγ2+C2s2\ne−(Ds2+3γ)t.(C22)\nSubstitutingtheinitialconditionsfortheradialspinhelix\n(1-3) into Eqs. (C15,C16) and performing integration\nwe find that the generalized functions ˜Sr(s) and˜Sz(s),\nwhich correspond to the initial conditions (1) and (3),\nact as follows\n/integraldisplay+∞\n0dssF(s)˜Sr(s) =−S0d\ndk/integraldisplayk\n0dskF(s)√\nk2−s2,(C23)\n/integraldisplay+∞\n0dssF(s)˜Sz(s) =S0d\ndk/integraldisplayk\n0dssF(s)√\nk2−s2,(C24)\nwhereF(s) is a smooth enough function.\nSubstituting Eqs. (C23,C24) into the general solution\n(C17,C18) and taking into account the expressions for\nthe Green functions (C19-C22), we obtain the explicit\nformulae for the solution of the drift-diffusion equations\n(C1-C2) for the initial conditions (1-3):\nSr(r,t) =−S0d\ndk/integraldisplayk\n0dsJ1(sr)√\nk2−s2/bracketleftBig\nkcosh/parenleftBig\nt/radicalbig\nγ2+C2s2/parenrightBig\n+/parenleftbig\nkγ+Cs2/parenrightbigsinh/parenleftBig\nt/radicalbig\nγ2+C2s2/parenrightBig\n/radicalbig\nγ2+C2s2\ne−(Ds2+3γ)t,(C25)\nSz(r,t) =S0d\ndk/integraldisplayk\n0dssJ0(sr)√\nk2−s2/bracketleftBig\ncosh/parenleftBig\nt/radicalbig\nγ2+C2s2/parenrightBig\n+(kC−γ)sinh/parenleftBig\nt/radicalbig\nγ2+C2s2/parenrightBig\n/radicalbig\nγ2+C2s2\ne−(Ds2+3γ)t.(C26)\nAppendix D: Relaxation of plain spin helix\nIn the case of the plane spin helix defined by the ini-\ntial conditions (4-6), the solution of the drift-diffusion9\nequations (B3-B5) has the same harmonical spatial de-\npendence at any moment of time. Therefore, we can seek\nthe solution of this system of equations in the form\nSx(x,t) =A(k,t)sin(kx),Sy(x,t) = 0,(D1)\nSz(x,t) =B(k,t)cos(kx).(D2)\nSubstituting expressions (D1-D2) into the system (B3-\nB5) we obtain the same system of ordinary differential\nequations (C5-C6) for the functions A(k,t) andB(k,t)\nas in the axially symmetric case. Eqs. (D1-D2) to-\ngether with Eqs. (C7-C9) determine solutions of the\ndrift-diffusionequations(B3-B5)withsuchaspecific har-\nmonic spatial dependence. The arbitrary amplitudes C+andC−are calculated from the initial conditions (Eqs.\n(4-6)). Finally, the solution of drift-diffusion equations\n(B3-B5) for the plane spin helix is given by\nSx(x,t) =−S0sin(kx)/bracketleftBig\ncosh/parenleftBig\nt/radicalbig\nγ2+C2k2/parenrightBig\n+ (Ck+γ)sinh/parenleftBig\nt/radicalbig\nγ2+C2k2/parenrightBig\n/radicalbig\nγ2+C2k2\ne−(Dk2+3γ)t,(D3)\nSz(x,t) =S0cos(kx)/bracketleftBig\ncosh/parenleftBig\nt/radicalbig\nγ2+C2k2/parenrightBig\n+(Ck−γ)sinh/parenleftBig\nt/radicalbig\nγ2+C2k2/parenrightBig\n/radicalbig\nγ2+C2k2\ne−(Dk2+3γ)t.(D4)\n∗pershin@physics.sc.edu\n1D.D.Awschalom, N.Samarth, andD.Loss, eds., Semicon-\nductor Spintronics and Quantum Computation (Springer-\nVerlag, 2002).\n2I. Zutic, J. Fabian, and S. Das Sarma, Rev. Mod. Phys.\n76, 323 (2004).\n3M. I. Dyakonov and V. I. Perel’, Sov. Phys. Solid State 13,\n3023 (1972).\n4M. I. Dyakonov and V. Y. Kachorovskii, Sov. Phys. Semi-\ncond.20, 110 (1986).\n5A. A. Kiselev and K. W. Kim, Phys. Rev. B 61, 13115\n(2000).\n6E. Y. Sherman, Appl. Phys. lett 82, 209 (2003).\n7M. Q. Weng, M. W. Wu, and Q. W. Shi, Phys. Rev. B 69,\n125310 (2004).\n8Y. V. Pershin and V. Privman, Phys. Rev. B 69, 073310\n(2004).\n9Y. V. Pershin, Phys. Rev. B 71, 155317 (2005).\n10L. Jiang, M. Weng, M. Wu, and J. Cheng, J. Appl. Phys.\n98, 113702 (2005).\n11B. A. Bernevig, J. Orenstein, and S.-C. Zhang, Phys. Rev.\nLett.97, 236601 (2006).\n12P. Schwab, M. Dzierzawa, C. Gorini, and R. Raimondi,Phys. Rev. B 74, 155316 (2006).\n13J. D. Koralek, C. P. Weber, J. Orenstein, B. A. Bernevig,\nS.-C. Zhang, S. Mack, and D. D. Awschalom, Nature 458,\n610 (2009).\n14P. Kleinert and V. V. Bryksin, Phys. Rev. B 79, 045317\n(2009).\n15M. Duckheim, D. L. Maslov, and D. Loss, Phys. Rev. B\n80, 235327 (2009).\n16I. V. Tokatly and E. Y. Sherman, Ann. Phys. 325, 1104\n(2010).\n17Y. Bychkov and E. Rashba, JETP Lett. 39, 78 (1984).\n18G. Dresselhaus, Phys. Rev. 100, 580 (1955).\n19C. P. Weber, J. Orenstein, B. A. Bernevig, S.-C. Zhang,\nJ. Stephens, and D. D. Awschalom, Phys. Rev. Lett. 98,\n076604 (2007).\n20S. Saikin, Y. Pershin, and V. Privman, IEE-Proc. Circ.\nDev. Syst. 152, 366 (2005).\n21A. R. Cameron, P. Riblet, and A. Miller, Phys. Rev. Lett.\n76, 4793 (1996).\n22F. Reif, Fundamentals of Statistical and Thermal Physics\n(McGraw-Hill, 1965).\n23Y. Pershin, Phys. E 23, 226 (2004)." }, { "title": "1811.07681v2.Correlation_effects_in_spin_models_in_the_presence_of_a_spin_bath.pdf", "content": "Correlation effects in spin models in the presence of a spin bath\nÁlvaro Gómez-León1, Tim Cox2and Philip Stamp2\n1Instituto de Ciencia de Materiales, CSIC, Cantoblanco, Madrid E-28049, Spain and\n2Department of Physics and Astronomy, University of British Columbia,\n6224 Agricultural Road, Vancouver, British Columbia, V6T 1Z1, Canada\n(Dated: May 24, 2022)\nWe analyze the effect of a bath of spins interacting with a spin system in terms of the equation of\nmotion technique. We show that this formalism can be used with general spin systems and baths,\nand discuss the concrete case of a Quantum Ising model longitudinally coupled to the bath. We\nshow how the uncorrelated solutions change when spin-spin correlations are included, the properties\nof the quasiparticle excitations and the effect of internal dynamics in the spin bath.\nI. INTRODUCTION\nSpins systems are among the most studied models in\nphysics, but their properties are also closely related with\nsystems in many other fields of science1–3. These mod-\nels are interesting because they display many emergent\nphenomena due to collective effects, and in addition, be-\ncause of their relevance in the development of quantum\ncomputers4. Most works on spin models mainly consider\nclosed systems and study the equilibrium phase diagram,\nthe criticalexponents, and morerecently, alsothe quench\ndynamics5,6. Less it is known about these models in the\npresence of environments, specially for those described\nby a collection of localized modes. These are known to\nproduce radically different effects than the usual oscil-\nlator bath in certain regimes7–9. As the coupling to lo-\ncalized spin environments is generally not weak30, the\ncalculations are generally non-perturbative and must be\ndone carefully, which implies that in the case of quan-\ntum models, the effect of correlations can become quite\nimportant.\nFor practical purposes, the study of spin models has\nrecently become more important due to the develop-\nment of the first quantum simulators and quantum\ncomputers10–12. In this case, environments are ubiqui-\ntous and its interaction with the quantum computer dur-\ning the computation time needs to be properly under-\nstood – e.g., the fate of the quantum critical point (QCP)\nduring the quantum annealing process, in the presence of\nan environment. In the case of quantum computers made\nof flux qubits, spin environments can be produced by\nparamagnetic impurities and nuclear isotopes in the sub-\nstrate, and even if they weakly couple, their effect in the\ndynamics can be non-trivial13. Also it has been shown\nthat for superconducting qubits dielectric loss can domi-\nnate, and it is produced by effective two level systems14.\nTo understand the physical effect of spin bath environ-\nments we study a spin model coupled to a set of localized\nspins. We use the equation of motion (EOM) technique\nfor the Green’s functions, and a decoupling scheme based\non a hierarchical expansion in terms of correlations. This\napproach transforms the EOM into a Dyson’s equation\nfor the Green’s function, with explicit non-perturbative\nexpressions for the self-energy, which include a full fre-quency and momentum dependence. Hence it can also\nbe used to study quantum dynamics of excited states,\nas already shown in15. The lowest order solution agrees\nwith mean-field (MF) treatments and can explain some\ngeneral features of the effect of a baths of spins, as the\nsolutions are solely characterized by the connectivity be-\ntween Ising spins and bath spins (i.e., the three differ-\nent coordination numbers that one can define for this\nmodel). Higher order corrections depend on the spe-\ncific details of the bath, however, we numerically solve\nthe self-consistent equations for some specific cases and\ndemonstrate that quantum correlations tend to suppress\nthe effect of the bath of spins, with respect to the MF\ncalculations.\nII. MODEL AND METHOD\nWe consider a model with a central system and a\nbath. The central system is made of a set of interact-\ning spins coupled to an external field ~B. Similarly, the\nbath is made of a set of spins as well, but we assume\nthat these are non-interacting, although they can be af-\nfected by an external field ~\u000131. The total Hamiltonian is\nH=HS+HB, whereHSandHBare given by (Greek\nindices specify the direction of the spin vector and latin\nindices label the different sites):\nHS=\u0000X\n\u0016;iB\u0016S\u0016\ni\u0000X\n\u0016X\ni;j>iV\u0016\ni;jS\u0016\niS\u0016\nj(1)\nHB=\u0000X\n\u0016;l\u0001\u0016I\u0016\nl+X\n\u0016X\ni;lA\u0016\ni;lS\u0016\niI\u0016\nl(2)\nHScorresponds to a general spin model coupled to an\nexternal field ~Band with arbitrary interaction V\u0016\ni;j, while\nHBcorresponds to the bath Hamiltonian which couples\nto a field~\u0001and interacts with the central system via\nA\u0016\ni;l. To study the properties of the system we define the\nnext double-time Green’s functions16:\nG\u000b;\f\nn;m(t;t0) =\u0000ihS\u000b\nn(t) ;O\f\nm(t0)i (3)\nY\u000b;\f\nn;m(t;t0) =\u0000ihI\u000b\nn(t) ;O\f\nm(t0)i (4)arXiv:1811.07681v2 [cond-mat.str-el] 20 Sep 20192\nwhere ;indicates that the Green’s functions may be a\ntime ordered, retarded or advanced. Also O\f\nmis some\ngeneral spin operator at site m(i.e., it can represent a\nsystem or a bath spin). The equation of motion for the\nGreen’s function is:\ni@tG\u000b;\f\nn;m =\u000e(t\u0000t0)h\u0002\nS\u000b\nn;O\f\nm\u0003\n\u0006i+i\u000f\u0016\u000b\u001aB\u0016G\u001a;\f\nn;m(5)\n+i\u000f\u0016\u000b\u001a0\n@X\ni6=nV\u0016\ni;nG\u0016\u001a;\f\nin;m\u0000X\nlA\u0016\nn;lY\u0016\u001a;\f\nln;m1\nA\nwhere [:::]\u0006refers to anti-commutator or commuta-\ntor, respectively17,18(i.e., to a fermionic or bosonic\nGreen’s function). We have defined the three-spin cor-\nrelatorsG\u0016\u001a;\f\nin;m(t;t0) =\u0000ihS\u0016\niS\u001a\nn;O\f\nmiandY\u0017\u001a;\f\nln;m(t;t0) =\n\u0000ihI\u0017\nlS\u001a\nn;O\f\nmi.\nAn important feature of many-body systems is the hi-\nerarchical structure of the EOM, where interaction terms\nproducehigher-orderGreen’sfunctions. Todealwiththis\nhierarchy, one needs to devise a method to decouple the\ninfinite set of equations into smaller blocks, which cor-\nrectly captures the regime of interest. For that purpose\nwe separate the Green’s functions into their uncorrelated\nand correlated parts: G\u0016\u001a;\f\nin;m =hS\u0016\niiG\u001a;\f\nn;m+hS\u001a\nniG\u0016;\f\ni;m+\nG\u0016\u001a;\f\nin;m(we will use calligraphic letters to denote the cor-\nrelated parts). This separation is completely general and\njust transforms the initial EOM for the two-point func-\ntion in two coupled equations, one for the two-point func-\ntions and one for the correlated parts. Physically, the\ntermshS\u0016\niiG\u001a;\f\nn;m+hS\u001a\nniG\u0016;\f\ni;mare similar to the Hartree\nand Fock contributions. They capture the average effect\nof all the other spins on the two-point function, while\nthe correlated part G\u0016\u001a;\f\nin;mcontains the corrections due tocorrelations with additional spins.\nOur approximation to close the system of equations\nwill assume that correlated parts scale in some partic-\nular way with the parameters of the system, such that\nthey can be neglected when they are small enough19,20.\nWe will organize terms in inverse powers of the coordina-\ntion number Z, whereZgenerally denotes any coordina-\ntion number required to describe the model (for the Ising\nmodel just one coordination number is required, while for\nthe examples below we show that three different coordi-\nnation numbers are needed. In this case the smallest\ncoordination number will dominate the scaling). Doing\nthis, it is possible to systematically include higher or-\nder corrections by simply adding higher order correlated\nparts.\nIt must be mentioned that the scaling of correlations\nconverges more slowly in 1D (it goes as \u0018Z\u00001, with\nZ= 2). However, it is still an interesting case, because in\nlow dimensional systems quantum corrections are usually\nenhanced. This facilitates the study of quantum correc-\ntionsduetothespinbath, andforthisreasonwewillcon-\nsider the 1D case for the numerical results. However one\nmustkeepinmindthattheequationsaregeneralforarbi-\ntrary dimension. Furthermore, we will show that includ-\ning the correlated parts in the equation of motion, allows\nto recover the exact quantum critical point in absence of\nthe spin bath, thus providing a good benchmark for our\nresults. For accurate resultsin low dimensional cases, the\npresent technique could also be combined with different\nfermionization or bosonization techniques6, and straight-\nforwardly applying the same Green’s function analysis.\nLet us apply this decoupling scheme to Eq.5 and\nFourier transform to frequency domain. One finds:\n!G\u000b;\f\nn;m =h\u0002\nS\u000b\nn;O\f\nm\u0003\n\u0006i+iX\n\u0016\u000f\u0016\u000b\u001a0\n@B\u0016+X\ni6=nV\u0016\ni;nhS\u0016\nii\u0000X\nlA\u0016\nn;lhI\u0016\nli1\nAG\u001a;\f\nn;m (6)\n+iX\n\u0016\u000f\u0016\u000b\u001a0\n@X\ni6=nV\u0016\ni;n\u0010\nhS\u001a\nniG\u0016;\f\ni;m+G\u0016\u001a;\f\nin;m\u0011\n\u0000X\nlA\u0016\nn;l\u0010\nhS\u001a\nniY\u0016;\f\nl;m+Y\u0016\u001a;\f\nln;m\u00111\nA\nThe first line corresponds to the uncorrelated contribution, while the second line corresponds to the contribution from\ncorrelations between spins. Notice the appearance of the Green’s function Y\u0016;\f\nl;m=\u0000ihI\u0016\nl;O\f\nmi, which couples the\nequation of motion for the Ising spins with the one for the bath spins:\n!Y\u000b;\f\nn;m =h\u0002\nI\u000b\nn;O\f\nm\u0003\n\u0006i+iX\n\u0016\u000f\u0016\u000b\u001a \n\u0001\u0016\u0000X\niA\u0016\ni;nhS\u0016\nii!\nY\u001a;\f\nn;m (7)\n\u0000iX\n\u0016;i\u000f\u0016\u000b\u001aA\u0016\ni;n\u0010\nhI\u001a\nniG\u0016;\f\ni;m+Y\u001a\u0016;\f\nni;m\u0011\nFinally, one can write the equations in a more compact form using matrix notation:\n\u0010\n!\u0000^H\u0011\n\u0001^G= ^\u001f\u0006+^V\u0001^G (8)3\nwhere ^Gis a vector containing all the system and bath\ntwo-point functions, ^\u001f\u0006contains the source terms com-\ning from the commutators/anti-commutators, ^Gcontains\nall extra contributions from correlations, ^Vis the general\ninteraction matrix for the correlated parts, and ^His the\neffectiveHamiltonianforthetwo-pointGreen’sfunctions.\nIf the contribution from the correlated parts ^Gis small\nenough, the solution can be obtained by direct matrix in-\nversion. This is a good approximation near fixed points\nof the RG flow, where solutions are approximately de-\nscribed by free spin Hamiltonians with renormalized pa-\nrameters. Furthermore, if one defines ^g=\u0010\n!\u0000^H\u0011\u00001\n,\nthe previous equation resembles a Dyson’s equation, and\nfrom the correlated parts it can be shown that ^V\u0001^Gcan\nbe written as ^\u0006 (!)\u0001^G, giving rise to the self-energy and\nthe familiar Dyson’s equation.\n^G= ^g\u0001^\u001f\u0006+ ^g\u0001^\u0006\u0001^G (9)\nIII. UNCORRELATED SOLUTION\nTo close the system of equations one can neglect the\ncorrelated parts in Eq.8. Then the solution for the\nGreen’s function is given by a matrix inversion. Eqs.6\nand 7 have two contributions from each interaction term,\nsimilar to the well known Hartree and Fock terms. If\none ignores the Fock contribution, the coupling between\nGreen’sfunctionsatdifferentsitesvanishesandtheequa-\ntions are diagonal in real space. This is equivalent to\nderive an effective-medium Hamiltonian for each spin,\nwhere all correlations between sites are neglected. In this\ncase one finds poles at !s(n) =q\n(P\n\u000bh\u000bs(n))2for the\ncentral system, and at !b(n) =q\n(P\n\u000bh\u000b\nb(n))2for the\nbath, where h\u0016\ns(n) =B\u0016+P\ni6=nV\u0016\nn;iM\u0016\ni\u0000P\nlA\u0016\nn;lm\u0016\nl\nandh\u0016\nb(n) = \u0001\u0016\u0000P\niA\u0016\ni;nM\u0016\ni. As the Green’s functions\ndependonthelocalmagnetization M\u000b\nn=hS\u000b\nniandm\u000b\nn=\nhI\u000b\nni, they need to be determined self-consistency. Phys-\nically, this happens due to the non-linearities introduced\nby the interaction term. Defining the spectral function\nJ\u000b;\f(!) =i\u0002\ng\u000b;\f\nn;n(!+i\u000f)\u0000g\u000b;\f\nn;n(!\u0000i\u000f)\u0003\u0000\ne!\nT\u00061\u0001\u00001,\nthe self-consistency equations are obtained from the re-\nlation between the statistical average and the spectral\nfunctioni\n2\u000f\u0016\u0017\u000bM\u000b\nn=\u0001\nJ\u0017;\u0016(!)d!=2\u001916(the\u0006sign cor-\nresponds to fermionic or bosonic Green’s function, re-\nspectively). The previous solution for the Green’s func-\ntions, in combination with the solutions of these non-\nlinear equations characterizing the local magnetization\nM\u000b\nnandm\u000b\nl, fully determine the properties of the sys-\ntem if correlations can be neglected. Furthermore, as\nin this uncorrelated case the Green’s functions display\nsimple poles only, one can rewrite the integral over fre-\nquency as a sum over poles. Then the final form of theself-consistency equations is:\nM\u000b\nn=h\u000b\ns(n)\n2!s(n)tanh\u0012!s(n)\n2T\u0013\n(10)\nm\u000b\nn=h\u000b\nb(n)\n2!b(n)tanh\u0012!b(n)\n2T\u0013\n(11)\nThis result has been particularized for spin 1/2, but\nlarger spins can be studied similarly, as we discuss be-\nlow for the specific case of the Quantum Ising model.\nThis simple result is a good first estimate to see if the\nsolutions are characterized by a magnetic texture. For\nexample, it can capture the anti-ferromagnetic ordering\nfor the case Vi;j<0. With this information one can build\nmore complete solutions, based on the symmetries of the\nground states.\nQuantum Ising model longitudinally coupled to the spin\nbath:As a more specific case, let us consider the fer-\nromagnetic Quantum Ising model, longitudinally cou-\npled to a bath of spins (we set Vx;y\ni;j= 0,By;z= 0,\nAx;y\ni;l= 0and initially ~\u0001 = 0). In this case the poles for\nthe central system are !s=q\nB2+ (V0Mz\u0000ZBAmz)2,\nwhile the ones for the bath are !b=ZBSAMz. The\nself-consistency equations can be directly obtained from\nEq.10 (to simplify notation we rename V0=ZSV,Vz\ni;j=\nVandA=Az\ni;l). We have assumed homogeneous mag-\nnetization to simplify the expressions, which holds if the\nsystem has a uniform magnetization solution. This not\nobvious in the presence of the spin bath, as it can me-\ndiate interactions between the system spins, modulating\nthe initially spatially homogeneous solution. However,\nwe assume bath configurations which do not destroy the\nhomogeneous magnetization.\nWhen correlations are neglected, the physics is dom-\ninated by three coordination numbers or connectivities:\nZScorresponds to the number of spins which directly in-\nteract in the central system, ZBto the number of bath\nspins connected to each spin, and ZBSto the number\nof spins directly interacting with each bath spin. The\ncoordination number ZSis well known in the analysis\nof the Ising model, but with the addition of the bath\nof spins, now two additional coordination numbers are\nneeded. Interestingly, when correlations are neglected,\ntheresultsonlydependonthetopologyofthelattice(i.e.,\nthe graph connectivity), while the geometrical effects will\nappear when correlations between sites are added.\nAn obvious question now is how the different phase\ntransitions are affected by the bath. To obtain the Curie\ntemperature, Tcwe fixB= 0and expand to third order\nin powers of Mz. The solution for Mz= 0is found to be:\nTc'V0\n8+r\n\u0000V0\n2\u00012+ZBZBSA21+8j~Pj+4j~Pj2\n6\n4(12)\nIt is easy to see that Tc(A!0) =V0=4, in agreement\nwith the Curie-Weiss law for the Ising model. The crit-\nical temperature seems to increase with all coordination4\nnumbers and with the total spin of the bath\f\f\f~P\f\f\f(see de-\ntailsofthecalculationintheAppendixA,wherearbitrary\nspin value for the bath is assumed). This result requires\nsome clarification, as it is known that the T-dependence\nfor the longitudinal magnetization should not be affected\nif the spin couples to a single bath spin21. The reason for\nthis discrepancy is the lack of correlations between the\nIsing spin and the local bath spin for the uncorrelated\nsolution. If one considers instead the electro-nuclear ba-\nsis at each site, or includes correlations between the bath\nandtheIsingspin, bothresultsagree. Toconfirmthis, we\nhave calculated in the Appendix the deviation from the\nexact solution for the case of a single Ising spin coupled\nto a bath spin. It shows that the limits of high and low\ntemperature are well captured by the uncorrelated solu-\ntion. However, as one goes to intermediate temperatures,\ncorrelations between the two spins become important,\nand the exact solution deviates from the uncorrelated\none (Fig.5 in the Appendix). This provides a practical\nexample of the importance of correlations, specially near\nthe phase transitions, and also confirms the validity of\nour results near the fixed points of the theory.\nImportantly, our solution can describe the case in\nwhich the spin bath is non-local ( ZBS>1), which has\nnot been previously discussed. Also, as previously indi-\ncated, theelectro-nuclearcorrelationscanbeeasilyincor-\nporated by exactly diagonalizing the local electro-nuclear\nHamiltonian22.\nTo find the critical field Bcwhich characterizes the\nquantumphasetransition(QPT)wetakethelimit T!0\nand expand to first order in powers of Mz:\nMz'ZBA\f\f\f~P\f\f\f\n2r\nB2+A2\f\f\f~P\f\f\f2\nZ2\nB+B2V0Mz\n2\u0014\nB2+A2\f\f\f~P\f\f\f2\nZ2\nB\u00153=2(13)\nThis shows that Mz(T!0) = 0is not a solution, and\nthe QPT between the ferromagnetic and the paramag-\nnetic phases is blocked due to a remnant magnetization\ninduced by the bath. Furthermore, this remnant mag-\nnetization scales as the first term in Eq.13. This means,\nstrictly speaking, that there is no QPT, although trans-\nverse terms in the interaction can recover the QCP. In-\nterestingly, all three coordination numbers appear in the\nresult for the critical temperature, while ZBSis absent\nin the expression for the quantum critical point. Fig.1\ncompares the characteristic behavior of the longitudinal\nmagnetization in the presence of the spin bath for differ-\nent cases. It shows that at low temperatures, the longi-\ntudinal coupling to the bath blocks the QPT (red), but it\ncan be recovered by adding a transverse field \u0001x6= 0to\nthe bath spins (green). Increasing the temperature also\nunblocks the transition (blue), because it disorders the\nbath, but the phase transition is not strictly at T= 0.\nThe phase diagram in absence of correlations captures\nqualitatively the different phases, but the critical point is\noverestimated. For the 1D Ising model it is well known\nCorr.(A≠0)\nCorr.(A=0)\nRPA(A≠0)\nUncorr.(A≠0)\n0.0 0.5 1.0 1.5 2.00.00.10.20.30.40.5\nB/VMzFigure 1: Phase diagram for the uncorrelated solution as a\nfunctionof B=V. (Black)IsolatedIsingmodelatlowT.(Red)\nIsing model longitudinally coupled to a static bath for A=V =\n0:05andlow T. (Blue)Isingmodellongitudinallycoupledtoa\nstatic bath at higher T. Temperature disorders the spin bath\nandunblocksthephasetransition. (Green)Isingmodelatlow\nTlongitudinally coupled to a dynamical spin bath ( \u00016= 0),\nwhich unblocks the QPT as well. We have considered ZBS=\nZB= 1for the plot.\nthat the exact critical field corresponds to B=V = 1=2\n, and this discrepancy happens due to corrections pro-\nduced by correlations. In the last section we will show\nthataverygoodagreementisobtainedifcorrelatedparts\nare added, but let us first consider the effect of the Fock\nterms in the T= 0regime.\nIV. CORRELATIONS BETWEEN SPINS\nIn general, quantum systems are made of interacting\nparticles, and although their interactions may be weak,\nthe correlations generated by these interactions may still\nhave important consequences. For example, it is well\nknownthatnearclassicalandquantumphasetransitions,\nsystemsdisplaymacroscopiccorrelations, atlengthscales\nmuch larger than the typical length scales present in the\nmicroscopic model. This is a particular case of emer-\ngence, where the system behaves in a different way than\ntheparticlesintheunderlyingmicroscopicmodel. There-\nfore, it is clear that in some cases correlations are impor-\ntant, and results where correlations are neglected will be\nsignificantly affected. Another important case, present in\nlow dimensional quantum systems, is when due to con-\nfinement, the role of quantum fluctuations is enhanced.\nAwellknownconsequenceofthisistheabsenceofcertain\nphase transitions in low dimensional models.\nTo include the effect of correlations one just needs to\nkeep the terms that were previously neglected in the\nequation of motion. The simplest contribution is the\nFock term, which couples two-point Green’s functions at\ndifferent sites. With this term added, two-point corre-\nlations are captured. If the system is translationally in-\nvariant, the Green’s function is diagonal in k-space and5\nthe matrix inversion can be performed analytically. The\nmain effect of the Fock term is to make quasiparticles\ndispersive, and to interpolate between the fixed points\nof the theory, where correlations between sites can be\nneglected (e.g., between the B= 0and theV= 0lim-\nits of the Quantum Ising model). For the present case\nwith a spin bath, this correction can also correlate the\nIsing and the bath Green’s functions (see terms hS\u001a\nniY\u0016;\f\nl;m\nandhI\u001a\nniG\u0016;\f\ni;min Eq.6 and Eq.7, respectively). This im-\nplies that in general, magnons in the Ising model become\na mixture of Ising magnons and excitations in the spin\nbath. Finally, the correction due to the correlated part of\nthe three-point function requires to calculate a new equa-\ntion of motion, which in the present case will also encode\nmagnon-magnon interactions. Formally, it can be shown\nthat the equation of motion for the correlated part of the\nthree-point function can be written in the next form:\n\u0010\n!\u0000^HG\u0011\n\u0001^G=^\u00030+^\u0003\u0001^G (14)\nonce it is truncated by neglecting four-point correlations,\nwhich are expected to scale as Z\u0000232. In Eq.14 we have\ndefined HGas the “Hamiltonian” for the correlated part,\nwhich is obtained from its equation of motion, and ^\u0003\nand ^\u00030are the two types of source terms which can be\npresent. The solution can be written as:\n^G=^\u00030+^\u0003\u0001^G\n!\u0000^HG(15)\nInserting this result in Eq.8, yields the solution for the\ntwo-point function including correlations from the three-\npoint function:\n^G=^\u001f\u0006+^\u00060\n!\u0000^H\u0000^\u0006(16)\nwhere the self-energies are defined as ^\u0006\u0011^V\u0001^\u0003\n!\u0000HGand\n^\u00060\u0011^V\u0001^\u00030\n!\u0000HG. Equation 16 shows that in general, ^\u0006modi-\nfies the pole structure of the Green’s function , while ^\u00060\nmodifies the spectral weight.\nAlthough the formal description of the previous solu-\ntions seems quite simple, a full solution can be a chal-\nlenging task due to the self-consistency equations, and\nin general, one needs to make use of numerical methods.\nNevertheless, it is also possible to obtain some analyti-\ncalresultsbymeansofperturbativeexpansionsandother\napproximationmethods. Inthisworkwewillfocusonthe\neffect of two-spin correlations, which capture the magnon\nquasiparticles. The specific effect of the correlated parts\nwill be analyzed only for the correction to the longitu-\ndinal magnetization, while more general effects will be\ndiscussed in future works.\nCorrelations for the Quantum Ising model longitudi-\nnally coupled to the spin bath: Now we include correla-\ntions in the transverse Ising model, as the significance of\nthis model for current quantum computing architecturesiscrucial. TheadditionoftheFocktermmodifiesthepre-\nvious equations of motion, and can couple the Ising and\nbath spins. For the case with ~\u0001 = 0the main difference\nwiththeprevioussolutionisthechangeinthequasiparti-\ncle excitations, from localized spin flips to magnons with\ndispersion relation:\n!k=q\n(MzV0\u0000ZBAmz)2+B2\u0000BMxVk(17)\nNotice that the last term vanishes at both fixed points\nB= 0andV= 0, agreeing with the uncorrelated so-\nlutions. However, as one interpolates between the two,\nthe quasiparticles turn mobile and the band structure\nbecomes dispersive. Results including the Hartree and\nFock terms in the equations of motion are similar to\nthe Random Phase Approximation (RPA)22. The self-\nconsistency equations are now:\nMz=1\n2MzV0\u0000ZBAmz\n1\nNP\nk!kcoth\u0000!k\n2T\u0001 (18)\nMx=1\n2B\n1\nNP\nk!kcoth\u0000!k\n2T\u0001 (19)\nIn absence of the spin bath ( A= 0), the phase diagram\ndoes not change much with respect to the uncorrelated\nresult. The critical field Bcshifts towards slightly larger\nvalues, stabilizing the ferromagnetic phase. On the other\nhand, this approximation captures that, at the quantum\nphase transition, the magnon gap closes, as it can be seen\nin the spin-spin correlation function (Fig.2, top). When\nthe spin bath is included ( A6= 0), the difference between\nthe uncorrelated and correlated phase diagrams gets re-\nduced (see Fig.3, green and blue lines). The reason is\nthat two-point correlations between Ising spins dominate\nnear the unperturbed Ising critical point, which is now\nshifted. This can be seen in Fig.3, where the main dif-\nferences happen near the Ising critical point Bc'1:1V,\nand the two solutions match again for larger B. There-\nfore, spin-spin correlations do not change quantitatively\nthe phase diagram, in the presence of the spin bath. On\nthe other hand, the effect of the bath in the correlation\nfunctions is more crucial, as it leads to a mode soften-\ning at the critical point. Without the spin bath, the\nmagnon gap closes, signaling a divergence of the corre-\nlation length between Ising spins; however, the presence\nof the remnant field produced by the spin bath makes\nthe gap finite for all B, and the correlation length be-\ntween spins remains finite. Furthermore, the effect of a\ndynamical bath due to a transverse field \u00016= 0modi-\nfies this picture non-trivially. The fact that Eq.17 corre-\nsponds to Ising magnons under an effective longitudinal\nfield\u0018ZBAmzis not a coincidence. When the bath is\nstatic ( \u0001 = 0), system and bath do not get entangled,\nbut making the bath dynamical ( \u00016= 0) correlates both\nsystems and modifies the quasiparticle picture. Fig.4\nshows the appearance of an electro-nuclear mode with\nnon-vanishing spectral weight. This mode emerges near\n!'0when\u0014\u0011ZBZBSA2B\u0001mxMx6= 0and corre-\nsponds to a mixture of magnon and bath excitation, as6\nFigure 2: Complex part of the bosonic Green’s function\nGz;z\nk;k(!)as a function of kand!. (Top) In absence of the spin\nbath, the magnon gap closes at the critical point Bc'1:1V.\n(Bottom) Gap does not close at the critical point if the longi-\ntudinal coupling to the bath is included. We have considered\nA= 0:05Vand a 1D lattice with ZB=ZBS= 1. The results\nare expressed in logarithmic scale to enhance the contrast.\nCorr(A 0)\n0.00.51.01.52.00.00.10.20.30.40.5\nB/VMz\nFigure 3: MzvsB=Vfor the Ising model longitudinally cou-\npled to a bath of spins at T= 0. (Green) Uncorrelated solu-\ntion for A= 0:05VandZB= 1. (Blue) RPA solution includ-\ning for A= 0:05VandZB= 1. (Black) Solution including\ncorrelated parts without the spin bath ( A= 0). (Red) So-\nlution including correlations for a large spin bath ( A= 0:2V\nandZB= 100).\nit can be seen from the poles of the Green’s function:\n~!k=\u0006vuut\n2+!2\nk\n2\u0006s\n\u0014+\u0012\n2\u0000!2\nk\n2\u00132\n(20)\nFigure4: Emergenceofanelectro-nuclearmodeatthecritical\npoint when system and bath get correlated due to a finite\n\u0001 = 10\u00003B.\nwhere \n =p\n\u00012+Z2\nBSA2(details in the Appendix).\nThe presence of this mode has been observed in7and\ndiscussed in detail in ref.22. There it is shown that this\nmode does not carry spectral weight unless the bath is\ndynamical, andthatitallowstorecovertheQCP.Finally,\nwe discuss the impact of adding the correlated parts in\nthe equation of motion. As previously mentioned, their\neffect can be incorporated in terms of a self-energy, but\nin the general case, the self-energy will be a complicated\nfunction of frequency and momentum. To estimate their\neffect we analyze the longitudinal magnetization, which\ncan be addressed more easily by using the Majorana rep-\nresentation for spins23:\nS\u000b=\u0000i\n2\u000f\u000b\u00121\u00122\u0011\u00121\u0011\u00122(21)\nThis representation simplifies the self-consistency equa-\ntions, andmakestheself-energyforthelongitudinalmag-\nnetization \u0006zz\u0010\n~k;!\u0011\na function of !only. Furthermore,\nunlike the Wigner-Jordan transformation, this represen-\ntation is local and the separation between correlated and\nuncorrelated parts is analogous to the spins case. The\nderivation of the equations of motion, as well as the de-\ncoupling scheme is analogous to the case of the spin rep-\nresentation, as described in the Appendix.\nThe phase diagram including the correlated parts is\nshown in Fig.3. In absence of the spin bath ( A= 0,\nblackline), theferromagneticphase shrinksandthe QCP\nshifts toB=V'1=2, which is the exact result for the 1D\nIsing model. When the spin bath is included ( A6= 0,\nred line), the phase boundary shifts towards larger val-\nues ofB=V, as it happened in the uncorrelated phase\ndiagram. However, in this case the remnant magnetic\nfield produced by the spin bath is highly renormalized\nby the correlated parts. This can be seen from the spe-\ncific values chosen for the bath in Fig.3 ( A= 0:2Vand\nZB= 100), which for the uncorrelated solution would\nproduce a phase boundary shifted towards much larger\nvalues ofB=V. Importantly, although the QPT seems to\nbe recovered when correlated parts are included, this is7\nnot the case. One can check that Mz(T= 0) = 0 is not a\nfixed point of the self-consistency equations, although its\nfinal value is very small due to the large renormalization.\nConclusions: We have shown that a large variety of\nspin models interacting with spin baths can be treated\nusing the equation of motion technique. This approach\nallows for non-perturbative solutions, even to lowest or-\nder, and using our decoupling scheme one can include\ncorrections in a systematic way. Neglecting correlations\nyields the standard MF solution, while adding the lowest\norder correlations between pairs of spins reproduces the\nRPA results.\nAs a concrete example, we have analyzed the Quantum\nIsing model coupled to a local bath of spins, where each\nIsing spin is coupled to a set of ZBindependent spins.\nThis case is interesting due to the changes produced by\nthe longitudinal interaction on the critical properties. In\nthe absence of correlations we have obtained solutions\nfor arbitrary spin value in the bath, and demonstrated\nthat a remnant magnetic field produced by the bath can\nblock the quantum phase transition. When correlations\nare included, system and bath quasiparticles can mix,\ntransforming the standard magnons into electro-nuclear\nmodes. We have shown that the effect of internal dynam-\nics in the spin bath entangles the system and bath spins,and that this leads to the emergence of electro-nuclear\nmodes near the phase transition.\nWhen higher correlations are added (i.e., the corre-\nlated parts of the Green’s functions), some quantities are\naccurately captured. For the Quantum Ising model we\nhave obtained the exact critical point in 1D and shown\nthat the effect of the spin bath gets highly renormalized\nby the virtual processes.\nFuture directions for this work include the study of\nmorecomplicatedspinmodels, asforexamplewithtrans-\nverse terms and dipolar interactions, or considering more\ncomplex connectivity between bath and system (e.g., if\na bath spin can interact with more than a single Ising\nspin, theresultswouldappreciablychange). Toconclude,\neffective models of spin systems interacting with local-\nized modes are ubiquitous in experiments at low tem-\nperature. Some examples are: dangling bonds24, nuclear\nspins7,25,26, paramagnetic impurities27,28, and localized\nvibrational modes29. Therefore a a theoretical approach\nthat allows to systematically study them is very relevant\nfor the field.\nWe would like to thank R. D. McKenzie for helpful\ndiscussions. This work was supported by the National\nScientific and Engineering Research Council of Canada\nand A.G-L. acknowledges the Juan de la Cierva program.\n1E. B.-N. P.L. Krapivsky, S. Redner, A Kinetic View of\nStatistical Physics , Cambridge University Press, 2010.\n2S. Lloyd, J. Phys.: Conf. Ser. 302, 012037 (2011).\n3H. G. Hiscock et al., Proceedings of the National Academy\nof Sciences (2016).\n4A. Das and B. K. Chakrabarti, Rev. Mod. Phys. 80, 1061\n(2008).\n5K. Sengupta, S. Powell, and S. Sachdev, Phys. Rev. A 69,\n053616 (2004).\n6S. Sachdev, Quantum Phase Transitions , Cambridge Uni-\nversity Press, 2 edition, 2011.\n7H. M. Rønnow et al., Science 308, 389 (2005).\n8M. Schechter and P. C. E. Stamp, Phys. Rev. Lett. 95,\n267208 (2005).\n9N. V. Prokof’ev and P. C. E. Stamp, Reports on Progress\nin Physics 63, 669 (2000).\n10S. Boixo et al., Nature Physics 10, 218 (2014).\n11R. Barends et al., Nature 534, 222 (2016).\n12E. F. Dumitrescu et al., Phys. Rev. Lett. 120, 210501\n(2018).\n13S. Tomsovic, Proceedings from the Institute for Nuclear\nTheory-Vol.5: Tunneling in complex systems , World Sci-\nentific Pub Co Inc, 1998.\n14J. M. Martinis et al., Phys. Rev. Lett. 95, 210503 (2005).\n15C. Deng and X. Hu, Phys. Rev. B 73, 241303 (2006).\n16D. Zubarev, Soviet Physics Uspekhi 3, 320 (1960).\n17H.-Y. WANG, K.-Q. CHEN, and E.-G. WANG, Interna-\ntional Journal of Modern Physics B 16, 3803 (2002).\n18J. Schreiber, Physica B+C 96, 27 (1979).\n19A. Gómez-León, Phys. Rev. B 96, 064426 (2017).\n20A. Gómez-León, Phys. Rev. B 94, 035144 (2016).\n21J. H. Page, S. R. P. Smith, and D. R. Taylor, Journal ofPhysics C: Solid State Physics 17, 73 (1984).\n22R. D. McKenzie and P. C. E. Stamp, Phys. Rev. B 97,\n214430 (2018).\n23V. R. Vieira, Phys. Rev. B 23, 6043 (1981).\n24A.M.Holder, K.D.Osborn, C.Lobb, andC.B.Musgrave,\nPhysical review letters 111, 065901 (2013).\n25A. Morello, O. N. Bakharev, H. B. Brom, R. Sessoli, and\nL. J. de Jongh, Phys. Rev. Lett. 93, 197202 (2004).\n26E. Chekhovich et al., Nature materials 12, 494 (2013).\n27L. Luan et al., Scientific reports 5, 8119 (2015).\n28C. M. Quintana et al., Phys. Rev. Lett. 118, 057702\n(2017).\n29J. Lisenfeld et al., Scientific reports 6, 23786 (2016).\n30Typically oscillator bath couplings scale as 1=p\nNdue to\ntheir delocalized nature, being Nthe number of modes.\nTherefore perturbative calculations can, in most cases,\ncapture the main properties.\n31Interactions between bath spins can be easily introduced,\nbut it is known that large interactions would make the\nenvironment behave as an effective oscillator bath, as the\nmodes would not be localized anymore.\n32In the present case the system is characterized by three\ndifferent coordination numbers, and the smallest one will\ndominate this scaling.8\nAppendix A: Comparison between the uncorrelated and the exact solution for two spins\nHere we discuss the limitations of the uncorrelated solution in terms of a two spin system. Consider the next\nHamiltonian:\nH=\u0000BxSx\u0000BzSz\u0000\u0001zIz+AIzSz(A1)\nwhere a single spin is coupled to a single bath spin. The interaction between the two spins is longitudinal and\nproportional to A. In addition, the Ising spin has a longitudinal field Bzwhich mimics the effect of the field produced\nby all the other Ising spins in the Ising model, and a transverse field Bx. Also the bath spin can be biased by a\nlongitudinal field \u0001z. The exact calculation is possible due to the small size of the Hilbert space, and the statistical\naverage for the magnetization of the Ising spin can be obtained analytically:\nMz=Z\u00001Trn\ne\u0000H\nTSzo\n=e\u0000\u0001\n2T(A+ 2Bz)sinh\u0012p\n(A+2Bz)2+4B2\n4T\u0013\nZq\n(A+ 2Bz)2+ 4B2\u0000e\u0001\n2T(A\u00002Bz)sinh\u0012p\n(A\u00002Bz)2+4B2\n4T\u0013\nZq\n(A\u00002Bz)2+ 4B2(A2)\nwhere\nZ= 2e\u0001\n2Tcosh0\n@q\n(A\u00002Bz)2+ 4B2x\n4T1\nA+ 2e\u0000\u0001\n2Tcosh0\n@q\n(A+ 2Bz)2+ 4B2x\n4T1\nA (A3)\nis the partition function. On the other hand, we make use of the self-consistency equations for the magnetization,\nobtained from the uncorrelated solutions of the Green’s functions in the main text:\nMz=Bz\u0000Amz\n2q\nB2x+ (Bz\u0000Amz)2tanh0\n@q\nB2x+ (Bz\u0000Amz)2\n2T1\nA (A4)\nMx=Bx\n2q\nB2x+ (Bz\u0000Amz)2tanh0\n@q\nB2x+ (Bz\u0000Amz)2\n2T1\nA (A5)\nmz=\u0001z\u0000AMz\n2q\n(\u0001z\u0000AMz)2tanh0\n@q\n(\u0001z\u0000AMz)2\n2T1\nA (A6)\nThe comparison between the two solutions is shown in Fig.5. It shows that in the absence of a transverse field ( Bx'0)\nthe longitudinal bath does not affect the Curie temperature, which is missed by the uncorrelated solution. On the\nother hand, for non-vanishing transverse field (i.e., when the Ising model becomes “quantum”), the bath modifies the\nbehavior with temperature, specially in the low temperature regime.\nNotice that the exact solution in terms of Green’s functions is also possible in this case, by just including the corre-\nlated part. This indicates that the intermediate temperature regime needs to be characterized including correlations\n(as one would expect, specially at the Curie temperature, where the correlation length diverges).9\nA=10-3\nA=10-3(exact)\nA=5\nA=5(exact)\n0 1 2 3 40.00.10.20.30.4\nTMz\nA=10-3\nA=10-3(exact)\nA=5\nA=5(exact)\n0 1 2 3 40.00.10.20.30.40.5\nTMz\nFigure 5: Comparison between the self-consistent solution and the exact calculation. (Left) Case with Bz= 0:1andBx= 1,\nand (right) case with Bz= 1andBx= 0:1. In general, the uncorrelated solution and the exact one, perfectly agree at low and\nhigh temperature, while for intermediate values they differ (for weak Athey agree very well in general). When the system is\nin the analog of a FM phase, with Bx\u001cBz(right), the hyperfine coupling (even for large A) would not make an important\nchange in the Curie temperature (which happens for intermediate values of T). This is in contrast with the case Bx\u001dBz\n(left), where specially at low Tthe cases with small and large Aare quite different.\nAppendix B: Quantum Ising model coupled to a spin bath\nThe case of the Quantum Ising model is obtained from the general equation of motion by setting By;z= 0,Vx;y\ni;j= 0,\n\u0001x;y= 0andAx;y\ni;l= 0. Furthermore, if the bath spins are static ( ~\u0001 = 0), it is useful to consider for its basis the\nset of eigenstates jP;Pzi, wherePlabels the spin value P2h\f\f\f~P\f\f\f;\f\f\f~P\f\f\f\u00001;:::; 1=2i\nandPz2[P;P\u00001;:::;\u0000P]its\nprojection. Their equation of motion is calculated analogously, from the Heisenberg equation:\n@t^O=ih\nH;^Oi\n(B1)\nThe advantage of using this basis is that it allows to calculate the solutions for arbitrary spin value. The Hamiltonian\nreads in this basis:\nHS=\u0000BX\niSx\ni\u0000X\ni;j>iVi;jSz\niSz\nj\u0000\u0001zX\nlX\nP;PzPzXPz;Pz\nP;l+X\ni;lAi;lSz\niX\nP;PzPzXPz;Pz\nP;l(B2)\nwhere\nXPz;Pz\nP;l=jP;PzilhP;Pzjl\nis the projector onto the eigenstate of the bath at site l. Neglecting correlations one finds that the average value of\neach projector, and the magnetization mz\nn=P\nPz;PPzhXPz;Pz\nP;niare given by:\nhXa;a\nP;ni=e(a\u00001)hb\nT\n2\u0010\nehb\nT\u00001\u00112\ncoshh\nhb\nT\u0010\f\f\f~P\f\f\f+ 1\u0011i\n\u0000cosh\u0000hb\n2T\u0001 (B3)\nmz\nn=\u00001\n4csch\u0012hb\n2T\u00133 + cosh\u0000hb\nT\u0001\n\u00002\u0010\n2 +\f\f\f~P\f\f\f\u0011\ncoshh\nhb\n2T\u0010\n1 + 2\f\f\f~P\f\f\f\u0011i\n+ 2\f\f\f~P\f\f\fcoshh\nhb\n2T\u0010\n3 + 2\f\f\f~P\f\f\f\u0011i\ncosh\u0000hb\n2T\u0001\n\u0000coshh\nhb\nT\u0010\f\f\f~P\f\f\f+ 1\u0011i (B4)\nwherehb(n) = \u0001z\u0000P\niAi;nMz\ni. This solution is specific for half-integer spins, but the case for integer spins can be\ncalculated analogously. Finally, we use the solution for the fermionic Green’s function to obtain the self-consistency\nequations for the magnetization:\nMz=hs\n2!stanh\u0010!s\n2T\u0011\n(B5)\nMx=B\n2!stanh\u0010!s\n2T\u0011\n(B6)10\nWith these results one can now estimate the Curie temperature and the critical field. For the Curie temperature we\nare looking for solutions with B= 0. We consider the case where the bath is unbiased by an external field to simplify\nthe expressions ( \u0001z= 0). Expanding for small Mzup to third order, and solving for the value of Tthat makes\nMz= 0, we find:\nTc=V0\n8+1\n4vuut\u0012V0\n2\u00132\n+ZBZBSA21 + 8\f\f\f~P\f\f\f+ 4\f\f\f~P\f\f\f2\n6(B7)\nIt shows that for this approximation, the Curie temperature would increase with all the coordination numbers (re-\nmemberV0=ZsV), as well as with the total spin of the bath\f\f\f~P\f\f\f. As discussed in the previous section of the\nAppendix, this result is not correct for a local spin bath, because the Curie temperature should not be affected by\nthe longitudinally coupled bath. The correction is provided by the correlations between the Ising spins and the bath\nspins. On the other hand, this model can also have several Ising spins coupled to each bath spin, and in this case the\nCurie temperature can be affected. Similarly for the critical field one finds:\nMz'ZBA\f\f\f~P\f\f\f\n2r\nB2+A2\f\f\f~P\f\f\f2\nZ2\nB+B2V0Mz\n2\u0014\nB2+A2\f\f\f~P\f\f\f2\nZ2\nB\u00153=2+O\u0000\nM2\nz\u0001\n(B8)\nThenMz= 0is not a solution due to a remnant magnetization, which saturates for large A\u001dB.\nAppendix C: Effect of two-spin correlations\nIn order to include two-point correlations, we include the Fock term in the equation of motion. This couples Green’s\nfunctions at different lattice sites:\n!G\u000b;\f\nn;m =h\u0002\nS\u000b\nn;O\f\nm\u0003\n\u0006i+iX\n\u0016\u000f\u0016\u000b\u001a0\n@B\u0016+X\ni6=nV\u0016\ni;nM\u0016\ni\u0000X\nlA\u0016\nn;lm\u0016\nl1\nAG\u001a;\f\nn;m\n+iX\n\u0016\u000f\u0016\u000b\u001aM\u001a\nn0\n@X\ni6=nV\u0016\ni;nG\u0016;\f\ni;m\u0000X\nlA\u0016\nn;lY\u0016;\f\nl;m1\nA\n!Y\u000b;\f\nn;m =h\u0002\nI\u000b\nn;O\f\nm\u0003\n\u0006i+iX\n\u0016\u000f\u0016\u000b\u001a \n\u0001\u0016\u0000X\niA\u0016\ni;nhS\u0016\nii!\nY\u001a;\f\nn;m\u0000iX\n\u0016;i\u000f\u0016\u000b\u001aA\u0016\ni;nhI\u001a\nniG\u0016;\f\ni;m\nUsing a transformation to momentum space, while assuming spatially homogeneous solutions, we find:\n!G\u000b;\f\nk=h\u0002\nS\u000b;O\f\u0003\n\u0006i+iX\n\u0016\u000f\u0016\u000b\u001a!s;\u0016G\u001a;\f\nk+iX\n\u0016\u000f\u0016\u000b\u001aM\u001a\u0010\nV\u0016\nkG\u0016;\f\nk\u0000ZBA\u0016Y\u0016;\f\nk\u0011\n(C1)\n!Y\u000b;\f\nk=h\u0002\nI\u000b;O\f\u0003\n\u0006i+iX\n\u0016\u000f\u0016\u000b\u001a!b;\u0016Y\u001a;\f\nk\u0000iX\n\u0016\u000f\u0016\u000b\u001am\u001aZBSA\u0016G\u0016;\f\nk(C2)\nwhere!s;\u0016=B\u0016+P\ni6=nV\u0016\n0M\u0016\u0000ZBA\u0016m\u0016and!b;\u0016= \u0001\u0016\u0000ZBSA\u0016M\u0016.\nFor the case of the Ising model longitudinally coupled to a bath of spins and \u0001x= 0, the poles of the Ising spins\nGreen’s functions are at:\n!k=q\nB2+ (MzV0\u0000ZBAmz)2\u0000BMxVk (C3)\nThe corresponding self-consistency equations are obtained from the anti-commutator Green’s functions:\nMz=MzV0\u0000ZBAmz\n21\nNP\nk!kcoth\u0000!k\n2T\u0001 (C4)\nMx=B\n21\nNP\nk!kcoth\u0000!k\n2T\u0001 (C5)11\nIf the bath is static one can see that the bath occupation probabilities are unchanged and that Ising spins and bath\nspins are not entangled.\nIf the bath has internal dynamics, now driven by \u0001x6= 0, one finds that the poles are not simple Ising magnons\nanymore, but combinations of the bath and system excitations:\n!\u0006=vuut\n2+!2\nk\n2\u0006s\n\u0014+\u0012\n2\u0000!2\nk\n2\u00132\n(C6)\nwhere\u0014 =ZBZBSA2Bx\u0001xmxMx,!k=q\nB2x+ (Bz+V0Mz\u0000ZBAmz)2\u0000BxVkMxand \n =p\n\u00012x+ (\u0001z\u0000ZBSAMz). Analogously one can calculate the self-consistency equations for the magnetization,\nand the correlation functions:\nMz=!z\n2\u0018; Mx=Bx\n2\u0018(C7)\n\u0018\u00111\nNX\nk!+!\u0000\u0000\n!2\n+\u0000!2\n\u0000\u0001\n(1\u0000\u0015k)\u0002\u0000\n\n2\u0000!2\n\u0000\u0001\n!+tanh\u0000!\u0000\n2T\u0001\n\u0000\u0000\n\n2\u0000!2\n+\u0001\n!\u0000tanh\u0000!+\n2T\u0001\u0003\nwhere\n\u0015k=ZBA\u0001xMx\u0002\n!\u0000tanh\u0000!+\n2T\u0001\n\u0000!+tanh\u0000!\u0000\n2T\u0001\u0003\n\u0000\n\n2\u0000!2\n+\u0001\n!\u0000tanh\u0000!+\n2T\u0001\n\u0000\u0000\n\n2\u0000!2\n\u0000\u0001\n!+tanh\u0000!\u0000\n2T\u0001\n\u0002ZBSABxmx\u0002\n!\u0000tanh\u0000!+\n2T\u0001\n\u0000!+tanh\u0000!\u0000\n2T\u0001\u0003\n\u0000\n!2\nk\u0000!2\n+\u0001\n!\u0000tanh\u0000!+\n2T\u0001\n\u0000\u0000\n!2\nk\u0000!2\n\u0000\u0001\n!+tanh\u0000!\u0000\n2T\u0001\nTherefore the self-consistency equations can be written in an analogous form to the case ~\u0001 = 0, by just redefining\nthe function \u0018, which contains a mixing term proportional to \u0015k/ZBSZBA2Bx\u0001xMxmx. Importantly, in this case\nthe bath-bath correlators are also non-vanishing, as they get entangled through the Ising spins.\nAppendix D: Correlated parts\nThe equations of motion for the correlated parts are obtained analogously, by calculating the Heisenberg equations\nof motion:\n@tS\u001a\npS\u000b\nn=X\n\u0016;iB\u0016\u0000\n\u000f\u0016\u001a\u0012S\u0012\npS\u000b\nn+\u000f\u0016\u000b\u0012S\u001a\npS\u0012\nn\u0001\n+X\n\u0016X\ni;j>iV\u0016\ni;j\u0002\n\u000f\u0016\u001a\u0012\u0000\n\u000ej;pS\u0016\niS\u0012\np+\u000ei;pS\u0012\npS\u0016\nj\u0001\nS\u000b\nn+\u000f\u0016\u000b\u0012S\u001a\np\u0000\n\u000ej;nS\u0016\niS\u0012\nn+\u000ei;nS\u0012\nnS\u0016\nj\u0001\u0003\n\u0000X\n\u0016X\ni;lA\u0016\ni;lI\u0016\nl\u0000\n\u000ep;i\u000f\u0016\u001a\u0012S\u0012\npS\u000b\nn+\u000en;i\u000f\u0016\u000b\u0012S\u001a\npS\u0012\nn\u0001\n(D1)\n@tI\u001a\npS\u000b\nn=X\n\u0016;\u0012\u000f\u0016\u000b\u0012B\u0016I\u001a\npS\u0012\nn+\u000f\u0016\u000b\u0012X\n\u0016;\u0012X\ni;j>iV\u0016\ni;jI\u001a\np\u0000\n\u000ej;nS\u0016\niS\u0012\nn+\u000ei;nS\u0012\nnS\u0016\nj\u0001\n+X\n\u0016;\u0012\u000f\u0016\u001a\u0012\u0001\u0016I\u0012\npS\u000b\nn\n\u0000X\n\u0016X\ni;lA\u0016\ni;l\u0000\n\u000ei;n\u000f\u0016\u000b\u0012I\u0016\nlI\u001a\npS\u0012\nn+\u000el;p\u000f\u0016\u001a\u0012I\u0012\npS\u000b\nnS\u0016\ni\u0001\n(D2)\n@tI\u001a\npI\u000b\nn=X\n\u0016 \n\u0001\u0016\u0000X\niS\u0016\ni!\n\u0000\n\u000f\u0016\u001a\u0012A\u0016\ni;pI\u0012\npI\u000b\nn+\u000f\u0016\u000b\u0012A\u0016\ni;nI\u001a\npI\u0012\nn\u0001\n(D3)\nThe correlated part of the Green’s functions are obtained subtracting the uncorrelated contributions. For example for\nthe case ofG\u001a\u000b;\f\npn;mone hasi@tG\u001a\u000b;\f\npn;m =i@tG\u001a\u000b;\f\npn;m\u0000ihS\u001a\npi@tG\u000b;\f\nn;m\u0000ihS\u000b\nni@tG\u001a;\f\np;m. Finally, separating into uncorrelated12\nand correlated parts, and neglecting higher order correlators, one finds:\n!G\u001a\u000b;\f\npn;m' h\u0002\nS\u001a\npS\u000b\nn;O\f\nm\u0003\n\u0006i\u0000hS\u001a\npih\u0002\nS\u000b\nn;O\f\nm\u0003\n\u0006i\u0000hS\u000b\nnih\u0002\nS\u001a\np;O\f\nm\u0003\n\u0006i\n+iX\n\u0016;\u0012\u000f\u0016\u001a\u0012B\u0016G\u0012\u000b;\f\npn;m +iX\n\u0016;\u0012\u000f\u0016\u000b\u0012B\u0016G\u001a\u0012;\f\npn;m\n+iX\n\u0016\u000f\u0016\u001a\u0012X\ni6=p;nV\u0016\ni;p\u0010\nhS\u0016\niS\u000b\nnicG\u0012;\f\np;m+hS\u000b\nnS\u0012\npicG\u0016;\f\ni;m+hS\u0016\niiG\u0012\u000b;\f\npn;m +hS\u0012\npiG\u0016\u000b;\f\nin;m\u0011\n+iX\n\u0016\u000f\u0016\u000b\u0012X\ni6=p;nV\u0016\ni;n\u0010\nhS\u0016\niS\u001a\npicG\u0012;\f\nn;m+hS\u001a\npS\u0012\nnicG\u0016;\f\ni;m+hS\u0016\niiG\u001a\u0012;\f\npn;m +hS\u0012\nniG\u0016\u001a;\f\nip;m\u0011\n\u0000iX\n\u0016\u000f\u0016\u001a\u0012X\nlA\u0016\np;l\u0010\nhS\u000b\nnI\u0016\nlicG\u0012;\f\np;m+hS\u000b\nnS\u0012\npicY\u0016;\f\nl;m+hI\u0016\nliG\u0012\u000b;\f\npn;m +hS\u0012\npiY\u0016\u000b;\f\nln;m\u0011\n\u0000iX\n\u0016\u000f\u0016\u000b\u0012X\nlA\u0016\nn;l\u0010\nhS\u001a\npI\u0016\nlicG\u0012;\f\nn;m+hS\u001a\npS\u0012\nnicY\u0016;\f\nl;m+hI\u0016\nliG\u001a\u0012;\f\npn;m +hS\u0012\nniY\u0016\u001a;\f\nlp;m\u0011\n+i\n4X\n\u0012\u000f\u000b\u001a\u0012\u0000\nV\u000b\nn;pG\u0012;\f\np;m\u0000V\u001a\np;nG\u0012;\f\nn;m\u0001\n\u0000iX\n\u0016;\u0012\u000f\u001a\u0016\u0012V\u0012\nn;p\u0000\nhS\u000b\nnihS\u0016\npiG\u0012;\f\nn;m+hS\u000b\nnihS\u0012\nniG\u0016;\f\np;m+hS\u0012\nnihS\u0016\npiG\u000b;\f\nn;m\u0001\n+iX\n\u0016;\u0012\u000f\u000b\u0016\u0012V\u0016\np;n\u0000\nhS\u001a\npihS\u0016\npiG\u0012;\f\nn;m+hS\u001a\npihS\u0012\nniG\u0016;\f\np;m+hS\u0012\nnihS\u0016\npiG\u001a;\f\np;m\u0001\nSimilarly for the other Green’s functions:\n!Y\u001a\u000b;\f\npn;m' h\u0002\nI\u001a\npS\u000b\nn;O\f\nm\u0003\n\u0006i\u0000hI\u001a\npih\u0002\nS\u000b\nn;O\f\nm\u0003\n\u0006i\u0000hS\u000b\nnih\u0002\nI\u001a\np;O\f\nm\u0003\n\u0006i\n+iX\n\u0016;\u0012\u0000\n\u000f\u0016\u000b\u0012B\u0016Y\u001a\u0012;\f\npn;m +\u000f\u0016\u001a\u0012\u0001\u0016Y\u0012\u000b;\f\npn;m\u0001\n+iX\n\u0016;\u0012\u000f\u0016\u000b\u0012X\ni6=nV\u0016\nn;i\u0010\nhI\u001a\npS\u0016\niicG\u0012;\f\nn;m+hI\u001a\npS\u0012\nnicG\u0016;\f\ni;m+hS\u0012\nniY\u001a\u0016;\f\npi;m+hS\u0016\niiY\u001a\u0012;\f\npn;m\u0011\n\u0000iX\n\u0016;\u0012\u000f\u0016\u000b\u0012X\nl6=pA\u0016\nn;l\u0010\nhI\u001a\npI\u0016\nlicG\u0012;\f\nn;m+hI\u001a\npS\u0012\nnicY\u0016;\f\nl;m+hI\u0016\nliY\u001a\u0012;\f\npn;m +hS\u0012\nniW\u0016\u001a;\f\nlp;m\u0011\n\u0000iX\n\u0016;\u0012\u000f\u0016\u001a\u0012X\ni6=nA\u0016\ni;p\u0010\nhI\u0012\npS\u0016\niiG\u000b;\f\nn;m+hI\u0012\npS\u000b\nnicG\u0016;\f\ni;m+hS\u000b\nnS\u0016\niicY\u0012;\f\np;m+hI\u0012\npiG\u0016\u000b;\f\nin;m+hS\u0016\niiY\u0012\u000b;\f\npn;m\u0011\n\u0000iX\n\u0016;\u0012\u000f\u0016\u000b\u0012A\u0016\nn;p\u0002\u0000\nhI\u0016\npI\u001a\npi\u0000hI\u001a\npihI\u0016\npi\u0001\nG\u0012;\f\nn;m+hS\u0012\nni\u0000\nW\u0016\u001a;\f\npp;m\u0000hI\u001a\npiY\u0016;\f\np;m\u0001\u0003\n\u0000iX\n\u0016;\u0012\u000f\u0016\u001a\u0012A\u0016\nn;p\u0000\nhI\u0012\npiG\u000b\u0016;\f\nnn;m\u0000hS\u000b\nnihI\u001a\npiG\u0016;\f\nn;m+hS\u000b\nnS\u0016\nniY\u0012;\f\np;m\u0000hS\u000b\nnihS\u0016\nniY\u001a;\f\np;m\u0001\nThe two-bath spins correlation function is also needed:\n!W\u001a\u000b;\f\npn;m =h\u0002\nI\u001a\npI\u000b\nn;O\f\nm\u0003\n\u0006i\u0000hI\u001a\npih\u0002\nI\u000b\nn;O\f\nm\u0003\n\u0006i\u0000hI\u000b\nnih\u0002\nI\u001a\np;O\f\nm\u0003\n\u0006i\n+iX\n\u0016\u000f\u0016\u001a\u0012\"\n\u0001\u0016W\u0012\u000b;\f\npn;m\u0000X\niA\u0016\ni;p\u0010\nhI\u000b\nnI\u0012\npicG\u0016;\f\ni;m+hI\u000b\nnS\u0016\niicY\u0012;\f\np;m+hI\u0012\npiY\u000b\u0016;\f\nni;m +hS\u0016\niiW\u0012\u000b;\f\npn;m\u0011#\n+iX\n\u0016\u000f\u0016\u000b\u0012\"\n\u0001\u0016W\u001a\u0012;\f\npn;m\u0000X\niA\u0016\ni;n\u0010\nhI\u001a\npI\u0012\nnicG\u0016;\f\ni;m+hI\u001a\npS\u0016\niicY\u0012;\f\nn;m+hI\u0012\nniY\u001a\u0016;\f\npi;m+hS\u0016\niiW\u001a\u0012;\f\npn;m\u0011#\nThese set of equations of motion provides very complex expressions for the self-energy, and their role will be analyzed\nin future publications, however to estimate their contribution, we consider the specific case of the correction to the\nlongitudinal magnetization. Furthermore, we rewrite these equations of motion in terms of Majorana fermions to13\nsimplify the self-consistency equations:\nS\u000b\nn=\u0000i\n2\u000f\u000b\u00121\u00122\u0011\u00121\nn\u0011\u00122\nn (D4)\nI\u000b\nn=\u0000i\n2\u000f\u000b\u00121\u00122\r\u00121\nn\r\u00122\nn (D5)\n\u0011\u000b\nnbeing a Majorana fermion at site nthat fulfills the usual anti-commutation relation for fermions, and in addition\n\u00112= 1=2. Similarly the \r\u000b\nnrefer to the Majorana fermions for the bath spins. The Green’s functions required for the\ncalculation of the magnetization M\u000b\nn=\u0000i\n2\u000f\u000b\u00121\u00122h\u0011\u00121n\u0011\u00122niare:\nG\u000b;\f\nn;m(t;t0) =\u0000ih\u0011\u000b\nn(t) ;\u0011\f\nm(t0)i (D6)\nP\u000b;\f\nn;m(t;t0) =\u0000ih\r\u000b\nn(t) ;\r\f\nm(t0)i (D7)\nThe longitudinal magnetization is obtained from the diagonal two-point function, with equation of motion:\n!G\u000b;\f\nn;n=\u000e\u000b;\f+i\u000f\u0016\u000b\u0012h\u0016\ns(n)G\u0012;\f\nn;n (D8)\n+1\n2\u000f\u0016\u000b\u0012\u000f\u0017\u00121\u00122X\ni6=nV\u0016;\u0017\nn;iG\u00121\u00122\u0012;\f\niin;n\n\u00001\n2\u000f\u0016\u000b\u0012\u000f\u0017\u00121\u00122X\nlA\u0016;\u0017\nn;lY\u00121\u00122\u0012;\f\nlln;n\nwhich also requires the calculation of the bath spins Green’s function:\n!P\u000b;\f\nl;l=\u000e\u000b;\f+i\u000f\u0016\u000b\u0012h\u0016\nb(l)P\u0012;\f\nl;l(D9)\n\u00001\n2\u000f\u0016\u00121\u00122\u000f\u0017\u000b\u0012X\niA\u0016;\u0017\ni;lR\u00121\u00122\u0012;\f\niil;l\nwhere we have defined h\u0016\ns(n) =B\u0016+P\ni6=nV\u0016;\u0017\nn;iM\u0017\ni\u0000P\nlA\u0016;\u0017\nn;lm\u0017\nl,h\u0016\nb(l) =B\u0016\u0000P\niA\u0017;\u0016\ni;lM\u0017\ni, andY\u00121\u00122\u0012;\f\nlln;nandR\u00121\u00122\u0012;\f\niil;l\nare the correlated parts of the mixed-Green’s functions Y\u00121\u00122\u0012;\f\nlln;n=\u0000ih\r\u00123\nl\r\u00124\nl\u0011\u00121n;\u0011\f\nniandR\u00121\u00122\u0012;\f\niil;l=\u0000ih\u0011\u00121\ni\u0011\u00122\ni\r\u0012\nl;\r\f\nli,\nrespectively.\nTo lowest order one recovers the MF expressions presented in the first sections of the main text. In this case, as\nwe are calculating the diagonal Green’s function G\u000b;\f\nn;n, there is no contribution from a Fock term to lowest order, and\nthe contribution from magnons comes in from the correlated parts. Their equations of motion are:\n!G\u001b1\u001b2\u000b;\f\nppn;n =\u0000iX\n\u0016B\u0016\u0000\n\u000f\u0016\u0012\u001b1G\u0012\u001b2\u000b;\f\nppn;n +\u000f\u0016\u0012\u001b2G\u001b1\u0012\u000b;\f\nppn;n +\u000f\u0016\u0012\u000bG\u001b1\u001b2\u0012;\f\nppn;n\u0001\n+1\n2X\n\u0016;\u0017\u000f\u0017\u00121\u00122X\nkA\u0016;\u0017\np;k\u000f\u0016\u0012\u001b1\u0010\nh\r\u00121\nk\r\u00122\nkiG\u0012\u001b2\u000b;\f\nppn;n +h\u0011\u0012\np\u0011\u001b2\npiY\u00121\u00122\u000b;\f\nkkn;n\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0017\u00121\u00122X\nkA\u0016;\u0017\np;k\u000f\u0016\u0012\u001b2\u0010\nh\r\u00121\nk\r\u00122\nkiG\u001b1\u0012\u000b;\f\nppn;n +h\u0011\u001b1\np\u0011\u0012\npiY\u00121\u00122\u000b;\f\nkkn;n\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0017\u00121\u00122X\nkA\u0016;\u0017\nn;k\u000f\u0016\u0012\u000b\u0010\nh\r\u00121\nk\r\u00122\nk\u0011\u001b1\np\u0011\u001b2\npiCG\u0012;\f\nn;n+h\r\u00121\nk\r\u00122\nkiG\u001b1\u001b2\u0012;\f\nppn;n\u0011\n\u00001\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122X\ni6=p;nV\u0016;\u0017\ni;p\u000f\u0017\u0012\u001b1\u0010\nh\u0011\u00121\ni\u0011\u00122\niiG\u0012\u001b2\u000b;\f\nppn;n +h\u0011\u0012\np\u0011\u001b2\npiG\u00121\u00122\u000b;\f\niin;n\u0011\n\u00001\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122X\ni6=p;nV\u0016;\u0017\ni;p\u000f\u0017\u0012\u001b2\u0010\nh\u0011\u00121\ni\u0011\u00122\niiG\u001b1\u0012\u000b;\f\nppn;n +h\u0011\u001b1\np\u0011\u0012\npiG\u00121\u00122\u000b;\f\niin;n\u0011\n\u00001\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122X\ni6=p;nV\u0016;\u0017\ni;n\u000f\u0017\u0012\u000b\u0010\nh\u0011\u001b1\np\u0011\u001b2\np\u0011\u00121\ni\u0011\u00122\niiCG\u0012;\f\nn;n+h\u0011\u00121\ni\u0011\u00122\niiG\u001b1\u001b2\u0012;\f\nppn;n\u0011\n\u00001\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122V\u0016;\u0017\nn;p\u0000\n\u000f\u0017\u0012\u001b1h\u0011\u0012\np\u0011\u001b2\npi+\u000f\u0017\u0012\u001b2h\u0011\u001b1\np\u0011\u0012\npi\u0001\u0000\nG\u00121\u00122\u000b;\f\nnnn;n\u0000h\u0011\u00121\nn\u0011\u00122\nniG\u000b;\f\nn;n\u0001\n\u00001\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122\u000f\u0017\u0012\u000bV\u0016;\u0017\np;n\u0000\nh\u0011\u001b1\np\u0011\u001b2\np\u0011\u00121\np\u0011\u00122\npi\u0000h\u0011\u001b1\np\u0011\u001b2\npih\u0011\u00121\np\u0011\u00122\npi\u0001\nG\u0012;\f\nn;n14\n!Y\u001b1\u001b2\u000b;\f\nlln;n=\u0000iX\n\u0016B\u0016\u000f\u0016\u0012\u000bY\u001b1\u001b2\u0012;\f\nlln;n\u0000iX\n\u0016\u0001\u0016\u0010\n\u000f\u0016\u0012\u001b1Y\u0012\u001b2\u000b;\f\nlln;n+\u000f\u0016\u0012\u001b2Y\u001b1\u0012\u000b;\f\nlln;n\u0011\n\u00001\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122\u000f\u0017\u0012\u000bX\ni6=nV\u0016;\u0017\ni;n\u0010\nh\r\u001b1\nl\r\u001b2\nl\u0011\u00121\ni\u0011\u00122\niiCG\u0012;\f\nn;n+h\u0011\u00121\ni\u0011\u00122\niiY\u001b1\u001b2\u0012;\f\nlln;n\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0017\u00121\u00122\u000f\u0016\u0012\u000bX\nr6=lA\u0016;\u0017\nn;r\u0010\nh\r\u001b1\nl\r\u001b2\nl\r\u00121\nr\r\u00122\nriCG\u0012;\f\nn;n+h\r\u00121\nr\r\u00122\nriY\u001b1\u001b2\u0012;\f\nlln;n\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122X\ni6=nA\u0016;\u0017\ni;l\u000f\u0017\u0012\u001b1\u0010\nh\r\u0012\nl\r\u001b2\nliG\u00121\u00122\u000b;\f\niin;n +h\u0011\u00121\ni\u0011\u00122\niiY\u0012\u001b2\u000b;\f\nlln;n\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122X\ni6=nA\u0016;\u0017\ni;l\u000f\u0017\u0012\u001b2\u0010\nh\r\u001b1\nl\r\u0012\nliG\u00121\u00122\u000b;\f\niin;n +h\u0011\u00121\ni\u0011\u00122\niiY\u001b1\u0012\u000b;\f\nlln;n\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122A\u0016;\u0017\nn;l\u0000\n\u000f\u0017\u0012\u001b1h\r\u0012\nl\r\u001b2\nli+\u000f\u0017\u0012\u001b2h\r\u001b1\nl\r\u0012\nli\u0001\nG\u00121\u00122\u000b;\f\nnnn;n\n+1\n2X\n\u0016;\u0017\u000f\u0017\u00121\u00122\u000f\u0016\u0012\u000bA\u0016;\u0017\nn;l\u0010\nh\r\u001b1\nl\r\u001b2\nl\r\u00121\nl\r\u00122\nli\u0000h\r\u001b1\nl\r\u001b2\nlih\r\u00121\nl\r\u00122\nli\u0011\nG\u0012;\f\nn;n\n!R\u001b1\u001b2\u000b;\f\nppl;l=\u0000iX\n\u0016B\u0016\u0010\n\u000f\u0016\u0012\u001b1R\u0012\u001b2\u000b;\f\nppl;l+\u000f\u0016\u0012\u001b2R\u001b1\u0012\u000b;\f\nppl;l\u0011\n\u0000iX\n\u0016\u0001\u0016\u000f\u0016\u0012\u000bR\u001b1\u001b2\u0012;\f\nppl;l\n\u00001\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122X\ni6=pV\u0016;\u0017\ni;p\u000f\u0017\u0012\u001b1\u0010\nh\u0011\u00121\ni\u0011\u00122\niiR\u0012\u001b2\u000b;\f\nppl;l+h\u0011\u0012\np\u0011\u001b2\npiR\u00121\u00122\u000b;\f\niil;l\u0011\n\u00001\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122X\ni6=pV\u0016;\u0017\ni;p\u000f\u0017\u0012\u001b2\u0010\nh\u0011\u00121\ni\u0011\u00122\niiR\u001b1\u0012\u000b;\f\nppl;l+h\u0011\u001b1\np\u0011\u0012\npiR\u00121\u00122\u000b;\f\niil;l\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122X\ni6=pA\u0016;\u0017\ni;l\u000f\u0017\u0012\u000b\u0010\nh\u0011\u001b1\np\u0011\u001b2\np\u0011\u00121\ni\u0011\u00122\niiCP\u0012;\f\nl;l+h\u0011\u00121\ni\u0011\u00122\niiR\u001b1\u001b2\u0012;\f\nppl;l\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0017\u00121\u00122X\nr6=lA\u0016;\u0017\np;r\u000f\u0016\u0012\u001b1\u0010\nh\u0011\u0012\np\u0011\u001b2\npiP\u00121\u00122\u000b;\f\nrrl;l+h\r\u00121\nr\r\u00122\nriR\u0012\u001b2\u000b;\f\nppl;l\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0017\u00121\u00122X\nr6=lA\u0016;\u0017\np;r\u000f\u0016\u0012\u001b2\u0010\nh\u0011\u001b1\np\u0011\u0012\npiP\u00121\u00122\u000b;\f\nrrl;l+h\r\u00121\nr\r\u00122\nriR\u001b1\u0012\u000b;\f\nppl;l\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122A\u0016;\u0017\np;l\u000f\u0017\u0012\u000b\u0000\nh\u0011\u00121\np\u0011\u00122\np\u0011\u001b1\np\u0011\u001b2\npi\u0000h\u0011\u001b1\np\u0011\u001b2\npih\u0011\u00121\np\u0011\u00122\npi\u0001\nP\u0012;\f\nl;l\n+1\n2X\n\u0016;\u0017\u000f\u0017\u00121\u00122A\u0016;\u0017\np;l\u0000\n\u000f\u0016\u0012\u001b1h\u0011\u0012\np\u0011\u001b2\npi+\u000f\u0016\u0012\u001b2h\u0011\u001b1\np\u0011\u0012\npi\u0001\u0010\nP\u000b\u00121\u00122;\f\nlll;l\u0000h\r\u00121\nl\r\u00122\nliP\u000b;\f\nl;l\u0011\n!P\u001b1\u001b2\u000b;\f\nrrl;l=\u0000iX\n\u0016\u0001\u0016\u0010\n\u000f\u0016\u0012\u001b1P\u0012\u001b2\u000b;\f\nrrl;l+\u000f\u0016\u0012\u001b2P\u001b1\u0012\u000b;\f\nrrl;l+\u000f\u0016\u0012\u000bP\u001b1\u001b2\u0012;\f\nrrl;l\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122X\niA\u0016;\u0017\ni;r\u000f\u0017\u0012\u001b1\u0010\nh\u0011\u00121\ni\u0011\u00122\niiP\u0012\u001b2\u000b;\f\nrrl;l+h\r\u0012\nr\r\u001b2\nriR\u00121\u00122\u000b;\f\niil;l\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122X\niA\u0016;\u0017\ni;r\u000f\u0017\u0012\u001b2\u0010\nh\u0011\u00121\ni\u0011\u00122\niiP\u001b1\u0012\u000b;\f\nrrl;l+h\r\u001b1\nr\r\u0012\nriR\u00121\u00122\u000b;\f\niil;l\u0011\n+1\n2X\n\u0016;\u0017\u000f\u0016\u00121\u00122\u000f\u0017\u0012\u000bX\niA\u0016;\u0017\ni;l\u0010\nh\u0011\u00121\ni\u0011\u00122\ni\r\u001b1\nr\r\u001b2\nriCP\u0012;\f\nl;l+h\u0011\u00121\ni\u0011\u00122\niiP\u001b1\u001b2\u0012;\f\nrrl;l\u0011\nwith the addition of the exact equation:\nGxyz;\f\nnnn;n (!) =i\n!(\u000ex;\fMx+\u000ey;\fMy+\u000ez;\fMz) (D10)15\nThese equations are exactly solved for the case of the quantum Ising model in the main text. Finally, we solve the\nself-consistency equations numerically, until full convergence. This produces the phase diagram shown in Fig.3." }, { "title": "2204.14087v1.Dynamics_of_a_Pair_of_Overlapping_Polar_Bright_Solitons_in_Spin_1_Bose_Einstein_Condensates.pdf", "content": "Dynamics of a Pair of Overlapping Polar Bright Solitons in Spin-1 Bose-Einstein Condensates\nGautam Hegde, Sandra M Jose, and Rejish Nath\nDepartment of Physics, Indian Institute of Science Education and Research, Dr. Homi Bhabha Road, Pune 411 008, India\nWe analyze the dynamics of both population and spin densities, emerging from the spatial overlap between\ntwo distinct polar bright solitons in Spin-1 Spinor Condensates. The dynamics of overlapping solitons in scalar\ncondensates exhibits soliton fusion, atomic switching from one soliton to another and repulsive dynamics de-\npending on the extent of overlap and the relative phase between the solitons. The scalar case also helps us\nunderstand the dynamics of the vector solitons. In the spinor case, non-trivial dynamics emerge in spatial and\nspin degrees of freedom. In the absence of spin changing collisions, we observe Josephson-like oscillations\nin the population dynamics of each spin component. In this case, the population dynamics is independent of\nthe relative phase, but the dynamics of the spin-density vector depends on it. The latter also witnesses the\nappearance of oscillating domain walls. The pair of overlapping polar solitons emerge as four ferromagnetic\nsolitons irrespective of the initial phase di \u000berence for identical spin-dependent and spin-independent interac-\ntion strengths. But the dynamics of final solitons depends explicitly on the relative phase. Depending on the\nratio of spin-dependent and spin-independent interaction strengths, a pair of oscillatons can also emerge as the\nfinal state. Then, increasing the extent of overlap may lead to the simultaneous formation of both a stationary\nferromagnetic soltion and a pair of oscillatons depending on the relative phase.\nI. INTRODUCTION\nBecause of the spin-dependent interactions, spinor conden-\nsates are ideal for exploring coherent spin-mixing dynamics\n[1–9]. The resulting oscillations in the populations of the Zee-\nman states can be engineered by varying the initial popula-\ntions and the relative phases [10]. In a spin-1 condensate, the\noscillatory dynamics arises because of the collisional inter-\nconversion of two atoms in the m=0 state and one atom each\ninm=1 and m=\u00001 states. Such a spin-changing process\npreserves the net magnetization. Further, high controllability\nover the spin-dynamics can be accessed via external magnetic\nor microwave fields utilizing linear and quadratic Zeeman ef-\nfects.\nSpinor condensates also provide an opportunity to study\nvector solitons [11–17] in quasi-one-dimensional (Q1D)\nBECs, including the bright solitons [18–21]. Vector soli-\ntons are self-trapped wave packets with multiple compo-\nnents. In spin-1 condensates, the bright solitons are classi-\nfied into polar and ferromagnetic ones based on the spin state\nand the expectation value of the time-reversal operator [21].\nCollisional properties of polar-polar, polar-ferromagnetic and\nferromagnetic-ferromagnetic solitons have been studied in the\npast [19, 22] and that later lead to the discovery of an exotic\nsoliton called the oscillaton [21, 23]. Oscillatons are solitons\nwhere the total density profile remains stationary while the\npopulations of each m-components oscillate in time. It is also\npossible to observe oscillatory spin-dynamics in a single soli-\nton [24], and in a pair of colliding solitons [23].\nIn this paper, we analyze the dynamics of a pair of overlap-\nping polar bright solitons in spin-1 spinor condensates. Note\nthat the scenario is di \u000berent from the setups used to study\nsoliton collisions. In the latter case, the solitons are initially\nplaced very far apart and collide against each other with an\ninitial velocity. The dynamics of overlapping optical solitons\nare previously studied both theoretically [25–27] and experi-\nmentally [28–31], but similar studies of matter-wave solitons\nare lacking. Because of that, first, we look at the dynamics\nof overlapping bright solitons in scalar condensates. The lat-ter helps us understand the unique features emerging from the\nvectorial nature of the spinor solitons. The dynamics critically\ndepends on the initial phase di \u000berence and the extent of over-\nlap between the solitons in the scalar case. It is expected that\nthe relative phase plays a vital role since it decides the nature\nof force between solitons [32]. Depending on the phase dif-\nference, the soliton fusion, atomic flow from one soliton to\nanother and repulsive dynamics can occur. Similar scenarios\nare found in the dynamics of overlapping optical solitons in\ndi\u000berent non-linear media [25, 28–30]. In particular, the flow\nof atoms from one soliton to another mimics the phenomenon\nof optical switching, and we term it atomic switching for the\nmatter-wave solitons.\nThe vectorial nature of spinor solitons leads to rich dynam-\nics, both in population and spin densities. Besides the rela-\ntive phase and extent of overlap between the solitons, the ratio\nbetween the spin-dependent and spin-independent interaction\nstrengths also a \u000bects the dynamics. A simplified picture of the\ndynamics is attained using a rotated frame. In the absence of\nspin changing collisions, we observe Josephson-like oscilla-\ntions in the population dynamics of each spin component. For\natoms in each Zeeman state, the Josephson-like oscillations\nare explained using an e \u000bective potential created by the den-\nsity of other components. In this case, the population dynam-\nics are independent of the relative phase, but that of the spin-\ndensity vector depends on it. The dynamics of spin-density\nvector reveals the appearance of oscillating domain walls, de-\npending on the relative phase. When spin-independent and\nspin-dependent interactions are identical, the collision of a\npair of polar solitons results in four ferromagnetic solitons ir-\nrespective of the value of the initial relative phase. Among the\nfour ferromagnetic solitons, each pair exhibits dynamics iden-\ntical to that of scalar solitons that become more apparent in\nthe new rotated frame. When the ratio of spin-dependent and\nspin-independent interactions is half, we see the formation of\na pair of oscillatons. An additional stationary ferromagnetic\nsoliton emerges if the extend of overlap is su \u000eciently large,\ndepending on the relative phase.\nThe paper is structured as follows. In Sec. II, we discussarXiv:2204.14087v1 [cond-mat.quant-gas] 29 Apr 20222\nFigure 1. Dynamics of two overlapping identical bright solitons\nin scalar condensates for di \u000berent relative phase between them for\ng=(2\u0019l2\n?)=\u00002. The separation \u0001 = 10l?for (a)-(c) and \u0001 = 5l?for\n(d)-(f). (a) and (d) for \u001e=0, and in (a) the overlap leads to perpetual\noscillating dynamics between the soliton fusion and the initial con-\nfiguration. In (d) the system did not reverse back completely to the\ninitial configuration. (b) and (e) for \u001e=\u0019=2, an initial atomic flow\nalong the phase gradient leads to asymmetric final solitons. (c) and\n(f) For\u001e=\u0019, the solitons repel each other, but identical in their size\nand shape.\nthe dynamics of overlapping bright solitons in scalar conden-\nsates. In Sec. III, we discuss the dynamics of overlapping po-\nlar solitons in spin-1 condensates. In particular, the Sec. III A\ndiscusses the physical setup of spinor solitons, especially the\ngoverning Hamiltonian, the initial state, time-dependent cou-\npled non-linear Schr ¨odinger equations in the original and the\nrotated frame. In Sec. III B, we classify the dynamics based on\nthe ratio between the spin-dependent and spin-independent in-\nteraction strengths. In particular, the dynamics in the absence\nof spin-changing collisions is discussed in Sec. III B 1. The\ndynamics for identical spin-independent and spin-dependent\ninteractions are discussed in Sec. III B 2 and when the spin-\ndependent interaction strength is half of the spin-independent\ninteraction is discussed in Sec. III B 3. Finally, we summarize\nin Sec. IV.\nII. SCALAR CONDENSATES\nIn the following, we analyze the dynamics of two identical,\noverlapping Q1D bright solitons in scalar condensates for dif-\nferent initial relative phases and overlaps. The analytic formof the initial two-soliton wavefunction is\n (z;t=0)=Ah\nsech (k(z\u0000\u0001=2))+ei\u001esech (k(z+ \u0001=2))i\n;\n(1)\nwhere\u001eis the relative phase, kis the wavenumber, \u0001is the ini-\ntial separation between the solitons, which controls the extent\nof overlap and the normalization constant is\nA=1\n2s\nk\n1+k\u0001cos\u001ecsch k\u0001:\nFor large values of \u0001, the solitons do not overlap, and they re-\nmain at rest, maintaining their size and shape over time. Inter-\nestingly, a tiny spatial overlap between the solitons can trigger\nnon-trivial dynamics, depending critically on the phase di \u000ber-\nence. The nature of soliton interactions, whether being attrac-\ntive or repulsive, also depends on their relative phase. The role\nof relative phase on the collisional dynamics of matter-wave\nsolitons has been a subject of study both experimentally [33]\nand theoretically [34]. Here, we analyze the dynamics mainly\nfor three di \u000berent relative phases: \u001e=0,\u001e=\u0019=2 rad and\n\u001e=\u0019rad, which captures the di \u000berent dynamical scenarios.\nFor the scalar condensates, the dynamics is governed by the\nQ1D non-linear Schr ¨odinger equation,\ni~@\n@t (z;t)=\"\n\u0000~2\n2M@2\n@z2+g\n2\u0019l?2j (z;t)j2#\n (z;t);(2)\nwhere g=4\u0019~2asN=Mwith the s-wave scattering length as<\n0,Nis the total number of particles and l?=p~=m!?is the\nwidth of the transverse harmonic confinement of frequency\n!?. The wavenumber kin Eq. (1) depends on the interaction\nstrength via k=jgj=(2\u0019~!?l4\n?).\nWhen\u001e=0, the overlap leads to the attractive interac-\ntion between the solitons, making them fuse into a single soli-\nton. Not being in its lowest energy state, the fused soliton\ndisentangles back into the initial two-soliton configuration if\nthe overlap is su \u000eciently small [see Fig. 1(a) for \u0001 = 10l?].\nThis process repeats periodically in time, leading to the per-\npetual oscillating dynamics between soliton fusion and the ini-\ntial configuration as shown in Fig. 1(a). As the initial overlap\nbetween the solitons increases (decreasing \u0001), the period of\noscillation gets shorter [see Fig. 2(a) for the time period Tvs\n\u0001]. Strikingly, oscillations in Fig. 1(a) arise in the absence of\nharmonic confinement, making it in high contrast with those\nexhibited by trapped solitons [33, 34]. Beyond a certain over-\nlap or smaller \u0001, the solitons become non-separable after the\ninitial fusion as shown in Fig. 1(d) for \u0001 = 5l?and the final\nsoliton exhibits breathing dynamics because of the extra en-\nergy it carries. These results are found to be similar to that of\noverlapping optical solitons in plasma [25].\nAs the initial relative phase between the solitons increases,\nthey repel each other. For instance, when \u001e=\u0019=2 [see\nFig. 1(b)], not only do the solitons repel, but also there is\na flow of atoms along the direction of phase gradient, i.e.,\nalong@\u001e=@ z, at the initial stage of the dynamics. Similar to\nthe atomic flow, an energy transfer among overlapping opti-\ncal solitons has been observed, leading to optical switching\n[28, 29]. In a similar spirit, the atom transfer from one soliton3\n7 8 910 11 12\n∆/l⊥306090120ω⊥T(a)\n7 8 910 11 12\n∆/l⊥0.00.10.20.3(b) ∆M/M (φ=π/2)\nvω⊥/l⊥(φ=π)\nFigure 2. The soliton properties as a function of \u0001for di \u000berent initial\nrelative phase for the scalar case. (a) The time period of oscillations\nbetween soliton fusion and the initial configuration for \u001e=0. In (b),\nthe solid line depicts the mass di \u000berence \u0001M=Mbetween the final\nsolitons for \u001e=\u0019=2 rad and the dashed line shows the speed of the\nfinal symmetric solitons for \u001e=\u0019rad. For both figures, g=(2\u0019l2\n?)=\n\u00002.\nto another can serve as a control knob to direct the motion of\nmatter waves. Hence we call it atomic switching. In Fig. 1(b),\nthe atomic flow takes place from the positive to negative z-\ndirection. Such a transient atomic current makes the final soli-\ntons asymmetric in density, size and speed. The denser soliton\nmoves faster. The speed of the lighter soliton arises from the\nrecoil of the atoms flown out from the original soliton. To\nquantify the mass asymmetry, we introduce the mass di \u000ber-\nence between the left and right solitons, i.e., \u0001M=Ml\u0000Mr,\nwhere Ml=MR0\n\u00001j (z;t)j2dzandMr=MR1\n0j (z;t)j2dz. In\nFig. 2(b), the solid line shows \u0001Mobtained when the solitons\nare well separated after a su \u000eciently long time. As seen, \u0001M\nincreases with decreasing \u0001because the larger the initial over-\nlap more significant the exchange of particles. The atomic\nswitching is evident in Figs. 1(b) and 1(e), respectively for\n\u0001 =10l?and\u0001 =5l?.\nThe above results imply that there is a net momentum in the\nsystem for\u001e=\u0019=2. To quantify that, we calculate the average\nmomentum of the initial wave function in Eq. (1) and is,\nhpi=4A2(1\u0000k\u0001cothk\u0001)csch k\u0001sin\u001e: (3)\nIn Fig. 3(a), we show hpias a function of \u001efor\u0001 = 10l?\nand\u0001 = 5l?, which oscillates between positive and negative\nvalues as a function of \u001e. The latter indicates that by tuning\nthe relative phase, we can control the direction of the transient\natomic current or the atomic switching. For \u001e=0,hpi=0\nas expected, and for \u001e=\u0019=2 rad, there is a non-vanishing\ninitial momentum leading to the asymmetric dynamics shown\nin Figs. 1(b) and 1(e). For \u001e=\u0019=2 rad,hpiexhibits a non-\nmonotonous behavior as a function of \u0001, exhibiting a local\nmaximum, as shown in Fig. 3(b). At \u0001 =0, the solitons com-\npletely overlap and remain at rest and thus hpi=0. Also, as\n\u0001!1 ,hpiapproaches zero as both solitons become com-\npletely independent.\nThe solitons also repel each other for \u001e=\u0019[see Figs. 1(c)\nand 1(f)], but there is no transient atomic current due to the\nnodal point at z=0. Hence, the final solitons are identical,\n−1 0 1\nφ/π−0.10.00.1/angbracketleftp/angbracketrightl⊥/¯h(a)\n∆ = 5l⊥\n∆ = 10l⊥\n0 10 20\n∆/l⊥0.000.040.080.120.16/angbracketleftp/angbracketrightl⊥/¯h(b)Figure 3. (a) The average momentum hpias a function of \u001efor\u0001 =\n5l?(solid line) and \u0001 = 10l?(dashed line) (b)hpivs\u0001for\u001e=\u0019=2\nrad.g=(2\u0019l3\n?~!?)=\u00002 for both (a) and (b).\nand\u0001M=0. The final solitons propagate with equal and\nopposite velocity, but the speed of each soliton depends on\nthe separation \u0001. The final speed vof the soliton vs \u0001for\n\u001e=\u0019rad is shown as a dashed line in Fig. 2(b). The more\noverlap the initial solitons have, the faster they move. Sum-\nmarizing this section, we see that the initial phase di \u000berence\nand the extent of overlap critically a \u000bect the dynamics of over-\nlapping bright solitons in scalar condensates. The most inter-\nesting feature is the local and transient atomic current in at-\ntractive overlapping condensates due to an e \u000bective repulsion\narising from the phase di \u000berence between the solitons. This\nscenario is identical to optical switching, o \u000bering the possi-\nbility of engineering matter-wave transport via controlling the\nrelative phase.\nIII. SPIN-1 SPINOR CONDENSATES\nA. Setup, model, initial state and rotated frame\nThe system we consider are partially overlapping spin-1\ncondensates, they are described by the Hamiltonian,\nˆH=Z\ndz2666664\u0000~2\n2MX\nmˆ y\nmd2\ndz2ˆ m+1\n2¯c0: ˆn2:+1\n2¯c1:ˆF2:3777775;\n(4)\nwhere Mis the mass of a boson, ¯ c0;1=c0;1=2\u0019l2\n?are the inter-\naction parameters, ˆ mis the field operator which annihilates\na boson of the mth Zeeman state, ˆ n(z)=Pf\nm=\u0000fˆ y\nm(z)ˆ m(z)\nis the total density operator. The components of spin density\noperator are\nˆF\u00172x;y;z(z)=X\nm;m0(f\u0017)mm0ˆ y\nm(z)ˆ m0(z); (5)\nwith f\u0017being the\u0017th component of the spin-1 matrices. The\nsymbol : : denotes the normal ordering that places annihi-\nlation operators to the right of the creation operators. The\nspin-independent and spin-dependent interaction parameters\narec0=(g0+2g2)=3 and c1=(g2\u0000g0)=3, respectively with\ngF=4\u0019~2aFN=mrelated to the scattering length aF=0;2of4\nz=0 zΔ m=1m=0m=-1\nFigure 4. Schematic setup of overlapping bright polar solitons in\nQ1D spin-1 condensates, along z-axis. A polar soliton with popu-\nlation shared among m=\u00061 states on the left (red-shaded) and a\npolar soliton with population solely at m=0 on the right side (blue-\nshaded). The separation \u0001determines the extent of overlap between\nthe two solitons.\nthe total spin-Fchannel. Nis the total number of atoms.\nSince we are interested in bright solitons, we always keep\nc0<0.\nWithin the mean-field theory, the dynamics of the sys-\ntem is described by the quasi-one-dimensional (Q1D) Gross-\nPitaevskii equations (GPEs) [18, 19, 23]\ni~@ 1\n@t=\"\n\u0000~2\n2M@2\n@z2+¯c0n+¯c1Fz#\n 1+c1p\n2F\u0000 0;(6)\ni~@ 0\n@t=\"\n\u0000~2\n2M@2\n@z2+¯c0n#\n 0+¯c1p\n2F+ 1+¯c1p\n2F\u0000 \u00001;(7)\ni~@ \u00001\n@t=\"\n\u0000~2\n2M@2\n@z2+¯c0n\u0000¯c1Fz#\n \u00001+¯c1p\n2F+ 0;(8)\nwhere n(z;t)=P\nmj m(z;t)j2is the total density, F\u0017(z;t)=P\nm;m0 \u0003\nm(f\u0017)mm0 m0, and F\u0006=Fx\u0006iFy. We introduce \r=\n\u0000c1=jc0jas the ratio of spin-dependent and spin-independent\ninteractions. The validity of Eqs. (6)-(8) requires that \u00161D\u001c\n~!?, where!?is the transverse confinement frequency and\n\u00161Dis the chemical potential of the Q1D condensates. We\nsolve Eqs. (6)-(8) numerically to analyze the dynamics [35].\nAt\r=1, Eqs. (6)-(8) represent a completely integrable\nsystem and support N-soliton solutions including two-soliton\nones [18].\nInitial state : The schematic diagram of the initial state is\nshown in Fig. 4, in which two distinct polar solitons overlap\naround z=0. The general solution of a static polar soliton is,\n (z)=r\nk\n2\u001fsech kz; (9)\nwhere k=j¯c0j=(4l?) is the inverse-width or the wavenumber\nof the soliton wavepacket, and \u001fis the spin state, which takes\nthe general form [21, 23],\n\u001f=ei\u001c0BBBBBBBBB@\u00001p\n2sin\u0012\ncos\u0012\n1p\n2ei\u001esin\u00121CCCCCCCCCA; (10)where\u001cis a global phase. The initial wave function of a pair\nof overlapping polar solitons we consider is,\n (z)=A\u0002sech (k(z+ \u0001=2))\u001fl+sech (k(z\u0000\u0001=2))\u001fr\u0003;(11)\nwhere A=p\nk=2 is the normalization constant, and \u001fr=\n(0;ei\u001e1;0)Tand\u001fl=(1;0;ei\u001e2)T=p\n2 are the spin states of the\nsolitons in right and left of z=0, respectively. Note that,\nindividually both polar solitons are degenerate ground state\nsolutions of the time-independent GPEs for ¯ c1>0. Compar-\ning to Eq. (10), \u0012=\u0019=2 for the left soliton and that of the\nright one is \u0012=0. The angles, \u001e1and\u001e2are the initial phases\nof the individual solitons and the extent of overlap between\nthe solitons is again controlled by \u0001. For Eq. (11),hpi=0,\nindependent of the value of \u0001and\u001e1.\nThe spin density vector is a null vector for polar soli-\ntons [18, 19, 21], but once they overlap, it may take a non-\nzero value in the overlapping region. For the initial state in\nEq. (11), the local spin density vector is,\nF(z;t=0)=4A2cos(\u001e1\u0000\u001e2=2)\ncosh 2 kz+cosh k\u0001\u0012\nˆxcos\u001e2\n2+ˆysin\u001e2\n2\u0013\n:\n(12)\nThe vector F(z;t=0) lies in the xy-plane and forms an an-\ngle\u001e2with the x-axis. The orientation is same at every point\nalong the z-axis, and depends only on \u001e2. The relative angle,\n\u001e1\u0000\u001e2=2 determines the magnitude of F(z;t=0) and hence,\nthe net magnetization. Integrating over z, we obtain the net\nspin density vector,\nFT=4A2\u0001cos(\u001e1\u0000\u001e2=2) csch k\u0001\u0012\nˆxcos\u001e2\n2+ˆysin\u001e2\n2\u0013\n;\n(13)\nwhich is a conserved quantity. Without loss of generality, we\nset\u001e2=0, which fixes the direction of FTalong the x-axis.\nIn the following we analyze the dynamics mainly for \u001e1=\n0; \u0019=2 and\u0019rad. For\u001e1=0 and\u001e1=\u0019rad, the initial\nspin-density vector is along the positive and negative x-axis,\nrespectively whereas for \u001e1=\u0019=2 rad, F(z;t=0) vanishes.\nIn the overlapping region, for \r,0, spin-mixing dynamics\ntakes place, in which two atoms in m=0 collisionally convert\ninto one atom each in m=1 and m=\u00001, and vice versa. As\nshown below, the nature of dynamics can be classified based\non the value of \rand\u001e1.\nRotated frame : Since FTis a conserved quantity, we can\nmove to a frame in which the quantization axis is parallel to\nthe direction of FT[24], which provides us with a simplified\npicture for the dynamics. Such a rotated frame is used in [21,\n23] to describe oscillatons, which emerged out of the collision\nbetween a ferromagnetic and a polar soliton. In our case, the\nnew frame is obtained by a rotation of \u0019=2 about the y-axis,\n(\u001f0=ei\u0019fy=2\u001f) i.e., making the quantization axis along the x-\naxis. The relation between the new and old spinor components5\nFigure 5. Dynamics of two overlapping polar bright solitons for \r=\n0,\u0001 = 10l?and ¯c0=~!?=\u00002. (a)-(c) show the dynamics of j 1j2+\nj \u00001j2,j 0j2and the total density n(z;t), respectively. The results are\nindependent of the value of \u001e1.\nare\n 0\n1(z)=1\n2 1(z)+1p\n2 0(z)+1\n2 \u00001(z) (14)\n 0\n0(z)=\u00001p\n2 1(z)+1p\n2 \u00001(z) (15)\n 0\n\u00001(z)= 1\n2 1(z)\u00001p\n2 0(z)+1\n2 \u00001(z)!\n: (16)\nNote that for the setup we consider 1and \u00001are identical\nat any instant, and hence 0\n0(z) completely vanishes. Hence,\nin the new frame, the system e \u000bectively reduces to a two-\ncomponent condensate [36]. The corresponding GPEs are,\ni~@ 0\n\u00061\n@t=\"\n\u0000~2\n2M@2\n@z2+¯c+j 0\n\u00061j2+¯c\u0000j 0\n\u00071j2#\n 0\n\u00061;(17)\nwhere ¯ c\u0006=¯c0\u0006¯c1. The exchange coupling between the two\ncomponents 0\n\u00061is provided by ¯ c0\u0000¯c1. When\r=1, ¯c0\u0000¯c1\nvanishes, indicating that in the rotated frame, the system ef-\nfectively reduces to two independent scalar condensates de-\nscribed by the wave functions 0\n1and 0\n\u00001. Note that 0\n1and\n 0\n\u00001, in general do not exhibit identical dynamics. We use the\nrotated frame to gain insights into the dynamics wherever pos-\nsible.\nB. Dynamics\nA comprehensive study of the dynamics of overlapping po-\nlar solitons as a function of \u0001,\rand\u001e1is a tedious task. We\nrestrict the analysis to \u001e1=0; \u0019=2; \u0019rad,\u0001 = 5l?;10l?and\n0\u0014\r\u00141 and in particular, (i) \r=0, (ii)\r=1 and (iii)\n\r=0:5, which capture the most exciting scenarios.\n−0.3−0.2−0.10.00.1Vm=1\neffω⊥t= 0 ω⊥t= 125 ω⊥t= 160\n−10−5 0 5 10\nz/l⊥−0.3−0.2−0.10.00.1Vm=0\neffFigure 6. E \u000bective potential Vm\ne f fgenerated for mth component by\nother components for \r=0,\u0001 = 10l?and ¯c0=~!?=\u00002. (a) Vm=1\ne f f\nexhibits a double well potential with its local minima and separation\nbetween them varying in time. (a) Vm=0\ne f foscillates in time between a\nsingle minimum and double minima. The results are independent of\nthe value of \u001e1\n1.\r=0\nFirst, we discuss the case for which there are no spin chang-\ning collisions, i.e., \r=0. In this case, the total population in\neach Zeeman component remains constant, and the population\ndynamics is found to be independent of \u001e1. We also found that\nchanging \u0001does not a \u000bect the dynamics qualitatively. The dy-\nnamics for\r=0 and \u0001 = 10l?is shown in Fig. 5. In partic-\nular, Figs. 5(a)-5(c) show the dynamics of j 1j2+j \u00001j2,j 0j2\nand the total density n(z;t), respectively. E \u000bectively, we ob-\nserve Josephson-like oscillations of populations in each com-\nponent [see Figs. 5(a) and 5(b)] between the two regions (left\nand right of z=0) where the solitons are initially placed.\nThe oscillations are non-sinusoidal. The density oscillations\nmay give a false impression that spin-changing collisions are\ntaking place. To demonstrate the Josephson-like oscillations,\nfor each Zeeman state, we introduce the population imbalance\nbetween the left and right sides of z=0, i.e.,\nZm(t)=1\nNm Z0\n\u00001dzj mj2\u0000Z1\n0dzj mj2!\n; (18)\nwhere Nmis the total number of atoms in mth state, which is a\nconstant.\nThe density oscillations of each component can be intu-\nitively understand via an e \u000bective potential created by other\ncomponents. First consider the population dynamics of m=1\nstate, which is shown in Fig. 5(a) (identical for m=\u00001 state\nalso). Since most of the population in m=1 is initially at\nthe left side of z=0, we have Z1(t=0)\u00181. For ¯ c0<0,6\n−101Zmm=±1 m= 0\n0 200 400 600 800 1000 1200\nω⊥t6810δ/l⊥\nFigure 7. Dynamics of (a) the population imbalance Zmand (b) the\nseparation\u000e(t) between the two peaks in the total density \r=0,\n\u0001 =10l?and ¯c0=~!?=\u00002. The results are independent of the value\nof\u001e1.\nthe terms ¯ c0j \u00001j2and ¯c0j 0j2in Eq. (6) form a double well\npotential for atoms occupying m=1 state, see Fig. 6(a) for\nVm=1\ne f f(z;t)=¯c0(j 0j2+j \u00001j2) at di \u000berent instants. A non-\nzero fraction of m=1 atoms in the right region triggers the\ntunnelling from left to right, leading eventually to oscillatory\ndynamics shown in Fig. 5(a). At t=0,Vm=1\ne f fis asymmetric\ninzsince m=0 population on the right side is twice that of\nm=\u00001 state on the left. As time evolves, Vm=1\ne f fgets mod-\nified, in particular, the two minima gets closer to each other\n[see dashed and dotted lines in Fig. 6(a)], which amplifies the\ntunnelling rate. Eventually the potential gets inverted but with\na shorter separation between the minima, as shown by dotted\nline in Fig. 6(a). The potential gets inverted because of the\nswapping of population in m=0 (m=\u00001) from left (right) to\nright (left). By this time, the majority of population in m=1\nstate has already tunnelled to the right. At this points, the re-\nverse dynamic happens, and the system recover to the initial\ndensity configuration. The whole processes then repeats peri-\nodically in time.\nAn identical picture also holds for the population dynamics\nof the m=\u00001 state. In contrast, the population in m=0, ini-\ntially placed in the right side, experiences a di \u000berent dynami-\ncal potential [ Vm=0\ne f f(z;t)=¯c0(j 1j2+j \u00001j2)] from the popula-\ntions of m=\u00061, as shown in Fig. 6(b). As time evolves, the\npotential minimum, initially on the left (solid line), moves to\nthe right via a transient double-well potential and then reverts.\nThem=0 atoms always move towards the potential mini-\nmum in the opposite region leading to the oscillatory dynam-\nics shown in Fig. 5(b). Also, note that m=\u00061 populations\nremain immiscible with the m=0 component for the entire\ntime, except in the overlapping region. The total density is\nFigure 8. Dynamics of spin-density Fx(z;t) for (a)\u001e1=0, (b)\u001e1=\n\u0019=2 rad, and (c) \u001e1=\u0019rad. Other parameters are \r=0,\u0001 = 10l?\nand ¯c0=~!?=\u00002.\ncharacterized by an out-of-phase oscillatory dynamics of the\ntwo peaks, see Fig. 5(c).\nIn Fig. 7(a), we show the dynamics of population imbalance\nZmfor the same dynamics shown in Fig. 5. Figure 7(b) shows\nthe separation ( \u000e) between the two peaks in the total density or\nequivalently the distance between the two minima in Vm=1\ne f f(z;t)\n[see Fig. 5(c)]. At t=0,\u000e= \u0001 and it varies periodically in\ntime. The population dynamics get faster as \u000edecreases and\nvice versa. It is seen in Fig. 7(a) that the population of each\ncomponent in the initial region never vanishes. Thus, there\nare always some atoms in m=\u00061 states on the left region and\nm=0 on the right region, which helps the system periodically\nrecover to initial densities.\nEven though the population dynamics shown in Fig. 5 is in-\ndependent of \u001e1, it a\u000bects the spin density vector F(z;t=0)=\nFx(z;t) ˆxvia Eq. (12) and its dynamics. In Figs. 8(a)-8(c), we\nshow the dynamics of Fx(z;t) for\u001e1=0,\u001e1=\u0019=2 rad, and\n\u001e1=\u0019rad, respectively. Since Fx(z;t)/( 1+ \u00001) \u0003\n0+\n( \u0003\n1+ \u0003\n\u00001) 0, the spin density vector is significant only in the\noverlapping regions. For \u001e1=0,FT=R1\n\u00001Fx(z;t)dzis pos-\nitive, i.e., the e \u000bective spin is pointing along the positive x-\ndirection. At t=0, the spin density vector is along the x-axis\nat every point in the overlapping region. As time progresses,\nwe observe the formation of domains with positive and nega-\ntiveFx. In particular, regions of negative values of Fxemerge\nfrom both the edges of the overlapping region, separated by a\nregion of positive Fx[see Fig. 8(a)]. The domain walls, which\nseparate the regions of positive and negative Fx, move towards\nthe centre, squeezing the region of positive Fxin the middle.\nThe latter results in an increase in spin density at the cen-\ntre. Because of the conservation of FTand energy, they can-\nnot reach beyond a separation. Eventually, the domain walls\nmove back to the edges, and their position oscillates in time.\nFor\u001e1=\u0019rad, the dynamics of F(z;t) is identical to that of\n\u001e1=0 except that positive and negative regions are switched\nas shown in Fig. 8(c).\nThe dynamics of F(z;t) for\u001e1=\u0019=2 rad is drastically dif-\nferent from that of \u001e1=0 and\u001e1=\u0019rad, see Fig. 8(b). For\n\u001e1=\u0019=2 rad, and F(z;0) vanishes and hence FT. As time7\nFigure 9. Dynamics of two overlapping polar bright solitons for \r=\n1,\u0001 = 10l?, ¯c0=~!?=\u00002, and\u001e1=0. (a)-(c) show the dynamics\nofj 1j2+j \u00001j2,j 0j2and the total density n(z;t), respectively. (d)-\n(f) show the dynamics of j 0\n1j2,j 0\n\u00001j2and the total density n0(z;t)=\nj 0\n1j2+j 0\n\u00001j2, respectively. Note that n(z;t)=n0(z;t).\nprogresses, we see simultaneous growth of equal regions of\npositive and negative Fx. The growth happens from the edges\nof the overlapping region, separated by a region of vanishing\nFx. As time evolves, the size of the central region of Fx=0\nshrinks and the regions of Fx,0 grow. After some time, sur-\nprisingly, Fx(z;t) vanishes quite rapidly and then reemerges\nwith opposite polarity. Then, the central region with Fx=0\ngrows until the regions of Fx,0 on either sides diminish.\nThe whole process repeats periodically in time and leads to\nthe butterfly pattern in space-time, as seen in Fig. 8(b). Thus,\nwe have a unique scenario of engineering the spin dynamics\nin spinor condensates leaving the population dynamics unaf-\nfected, by tuning the relative phase \u001e1, in the absence of spin-\nindependent interactions.\n2.\r=1\nThe presence of spin-changing collisions a \u000bects the dy-\nnamics drastically. For \r=1, spin-dependent and spin-\nindependent interactions are of equal strength. In Figs. 9(a)-\n9(c), we show the dynamics of j 1j2+j \u00001j2,j 0j2and the\ntotal density n(z;t), respectively for \r=1,\u0001 = 10l?and\n\u001e1=0. The first visible e \u000bect of spin-changing collision is\nthat all Zeeman components become spatially miscible, lead-\ning to identical density patterns for all Zeeman components at\nFigure 10. Dynamics of spin-density Fx(z;t) for (a)\u001e1=0, (b)\n\u001e1=\u0019=2 rad, and (c) \u001e1=\u0019rad. Other parameters are \r=1\nand\u0001 =10l?.\nlonger times. After a su \u000eciently long time, the initial, over-\nlapping polar solitons converted into four solitons via spin-\nchanging collisions, two each on either side of z=0. Each\nsoliton is characterized by a population ratio of 1 : 2 : 1 among\nthe Zeeman states ( m=\u00001,m=0,m=1) with a spin state\n\u001f=(1;\u0006p\n2;1)T=2, which is a ferromagnetic soliton [21, 23].\nThe spin density vector shown in Fig. 10(a) confirms that the\nfinal solitons are indeed ferromagnetic but it possesses oppo-\nsite directions for the inner and outer solitons. The inner soli-\ntons have Fx(z;t)>0, i.e., the polarization axis is along the\npositive x-axis, whereas the outer ones have the net spin vec-\ntor along the negative x-axis. Thus, we have a scenario where\ntwo overlapping polar solitons are dynamically converted into\nfour ferromagnetic solitons.\nIn the rotated frame [Eqs. (14)-(16)], the scenario becomes\nrelatively simple and also provide additional details on the dy-\nnamics. Recall that for \r=1, the wavefunctions 0\n1(z) and\n 0\n\u00001(z) are decoupled, and e \u000bectively we have two indepen-\ndent scalar condensates. At t=0, the initial states with \u001e1=0\nare,\n 0\n\u00061(z)=p\nk\n2\"sech (k(z\u0000\u0001=2))p\n2\u0006sech (k(z+ \u0001=2))p\n2#\n:(19)\nAs seen in Eq. (19), 0\n1and 0\n\u00001represent two-soliton states,\nidentical to Eq. (1), with relative phase \u001e=0 and\u001e=\u0019\nrad, respectively. Concurrently, comparing the results of\nthe original and the rotated frames in Fig. 9, we see that\nj 0\n1(z;t)j2provides the dynamics of the two inner solitons [see\nFig. 9(d)], andj 0\n\u00001(z;t)j2gives that of the two outer solitons\n[see Fig. 9(e)]. They both match with the dynamics of scalar\nsolitons shown in Fig. 1(a) and Fig. 1(c), respectively for the\ninner and outer solitons. In general, depending on the extend\nof the initial overlap, the inner and outer solitons have di \u000ber-\nent masses. The ratio of masses between the inner and outer\nsolitons can be easily obtained via the wavefunctions in the\nrotated frame as,\nR1\n\u00001j 0\n1j2dz\nR1\n\u00001j 0\n\u00001j2dz=1+k\u0001csch k\u0001\n1\u0000k\u0001csch k\u0001:8\nFigure 11. Dynamics of two overlapping polar bright solitons for\n\r=1,\u0001 = 10l?, ¯c0=~!?=\u00002, and\u001e1=\u0019=2 rad. (a)-(c) show the\ndynamics ofj 1j2+j \u00001j2,j 0j2and the total density n(z;t), respec-\ntively. (d)-(f) show the dynamics of j 0\n1j2,j 0\n\u00001j2and the total density\nn0(z;t)=j 0\n1j2+j 0\n\u00001j2, respectively. Note that n(z;t)=n0(z;t).\nHence, we can control the size of the inner and outer ferro-\nmagnetic solitons by varying \u0001.\nThe population dynamics for \u001e1=\u0019rad is identical to that\nof\u001e1=0 as discussed above, but the spin density of inner and\nouter solitons are now flipped, compare Figs. 10(a) and 10(c).\nFor\u001e1=\u0019=2 rad, population and spin density dynamics are\nqualitatively di \u000berent from the \u001e1=0 case. The population\ndynamics for \u001e1=\u0019=2 rad is shown in Fig. 11 and the corre-\nsponding spin density dynamics is shown in Fig. 10(b). After\nthe initial spin-mixing, the dynamics results in four ferromag-\nnetic solitons and they move away from the centre on either\nside. Due to the initial phase di \u000berence of\u0019=2, it takes a fi-\nnite time to reach a miscible state as seen in Figs. 11(a)-11(c),\nvia a transient oscillatory dynamics. The initial states in the\nrotated frame for \u001e1=\u0019=2 are\n 0\n\u00061(z)=p\nk\n2\"sech (k(z+ \u0001=2))p\n2\u0006isech (k(z\u0000\u0001=2))p\n2#\n(20)\nrepresenting two independent two soliton solutions, but with\nopposite phase gradients. As we know from the scalar case\nfor\u001e=\u0019=2 rad [see Figs. 1(b) and 1(e)], there is a transient\natomic current along the direction of phase gradient. There-\nfore, in 0\n1(z), there is a transient atomic current from left to\nright, and in 0\n\u00001(z), it is from right to left. That results in a\npair of asymmetric solitons in both 0\n1(z) and 0\n\u00001(z). Since\nFigure 12. Dynamics of two overlapping polar bright solitons for \r=\n0:5,\u0001 = 10l?, ¯c0=~!?=\u00002, and\u001e1=0. (a)-(c) show the dynamics\nofj 1j2+j \u00001j2,j 0j2and the total density n(z;t), respectively. (d)-\n(f) show the dynamics of j 0\n1j2,j 0\n\u00001j2and the total density n0(z;t)=\nj 0\n1j2+j 0\n\u00001j2, respectively. Note that n(z;t)=n0(z;t).\nthe denser solitons move faster, the inner solitons are lighter,\ncontrasting the \u001e1=0 case where all solitons are identical. As\nseen in Fig. 10(b), the spin vectors point in opposite directions\namong the inner solitons and the same among the outer ones.\nAgain, by tuning \u0001, we can control the mass ratio between the\ninner and outer solitons, also the frequency of transient oscil-\nlation seen in the initial stage of the dynamics in Figs. 11(a)\nand 11(b).\n3.\r=0:5\nIn the following, we analyze the dynamics for \r=0:5 and\nobserve the emergence of a pair of non-identical oscillatons\npropagating in opposite directions. Oscillatons are solitons\nin which the total density profile remains stationary while the\npopulations in di \u000berent spin components oscillate with a con-\nstant frequency [21, 23]. In Figs. 12(a)-12(c), we show the\ndynamics ofj 1j2+j \u00001j2,j 0j2and the total density n(z;t),\nrespectively for \r=0:5,\u0001 = 10l?and\u001e1=0. In the ini-\ntial stage, spin-mixing leads to oscillatory dynamics. Later,\nthe condensates transform into a pair of independent oscilla-\ntons moving in opposite directions. Unlike the case of \r=1\nwhere we see four final solitons, for \r=0:5, we see only\ntwo final solitons. It is explained later using the wavefunc-\ntions in the new frame. The initial oscillatory dynamics takes9\nFigure 13. (a) The snapshot of the initial state of two polar solitons.\n(b) The snapshot of densities at !?t=220 (c) The density profile\nof the left oscillaton at !?t=420. (d) The density profile of the\nright oscillaton at !?t=445. For all plots \r=0:5,\u0001 = 10l?,\n¯c0=~!?=\u00002, and\u001e1=0. The solid lines shows j 1j2+j \u00001j2, dashed\nlines showj 0j2and dotted-dashed lines show the total density n(z).\nplace between the two configurations shown in Figs. 13(a)\nand 13(b). Figure 13(b) is a completely miscible state with\npeaks ofj \u00061j2coincide with that of j 0j2, which implies that\nthe spin-mixing is taking place. On the oscillatory dynamics,\neach spin component leaves a small trail of atoms on the op-\nposite sides, i.e., m=\u00061 in the right region and m=0 in\nthe left region. Eventually, they cause the formation of a pair\nof oscillatons. Figures 13(c) and 13(d) are density snapshots\ntaken at two di \u000berent instants after the oscillatons are formed.\nThe total density n(z;t) of each oscillaton is identical in both\nfigures, whereas the densities of di \u000berent components vary in\ntime due to the spin-mixing. The total population oscillates\nbetween m=0 and m=\u00061 within each oscillaton. In general,\nthe frequency of internal oscillations of the two oscillatons are\ndi\u000berent and can be tuned via the extent of overlap, i.e. vary-\ning\u0001. Sincehpi=0, the denser oscillaton travels slower than\nthe other.\nBefore the oscillatons are formed, the whole system reaches\na miscible state. The density pattern of this miscible state is\nsusceptible to the initial noise in the system, if any present.\nThe latter can thereby a \u000bect the properties of the final oscil-\nlatons, such as the mass asymmetry, velocities and the fre-\nquency of the internal oscillations. But general qualitative\nfeatures of the dynamics remain the same. Such sensitivity\nto initial noise is not seen in any other cases discussed here.\nFor\u001e1=0, the initial states in the rotated frame are given\nby the two soliton solutions in Eq. (19). For \r=0:5, in the ro-\ntated frame, we have a binary condensate with attractive intra\nFigure 14. Density profiles of the oscillaton in the rotated frame\nat two di \u000berent instants. The solid ( j 0\n1(z)j2) and dashed (j 0\n\u00001(z)j2)\nlines show the the stationary solution obtained by solving Eqs. (21).\nThe dotted line shows the solutions from the dynamics. (a) for the\noscillaton in the left and (b) for the oscillaton in the right. Both snap\nshots are taken at !?t=1000.\nand inter-component interactions governed by the Eq. (17).\nTherefore both 0\n1(z) and 0\n\u00001(z) have the tendency to stick\ntogether in the same spatial regions, and leading to two fi-\nnal oscillatons or solitons. Note that, unlike in the lab frame,\nan oscillaton is characterized by stationary density profile for\neach component in the rotated frame, as seen in Figs. 12(d)-\n12(e). The wavefunctions for a stationary oscillation in the\nnew frame can be written as 0\n\u00061(z)=\u0011\u0006exp[i(\u0016\u0006t+\u001e\u0006)]\n[21, 23], where \u0011\u0006satisfy two coupled ordinary di \u000berential\nequations,\n \n\u0000~2\n2md2\ndz2+(¯c0+¯c1)\u00112\n\u0006+(¯c0\u0000¯c1)\u00112\n\u0007!\n\u0011\u0006=\u0016\u0006\u0011\u0006:(21)\nTaking the values of \u0016\u0006and\u001e\u0006from the numerical results\nand solving Eqs. (21) we obtain the stationary solutions \u0011\u0006.\nThe excellent agreement shown in Fig. 14 confirm us that\nthe final solitons formed in the dynamics are oscillatons. In\nFigs. 14(a) and 14(b), we show the density profiles of 0\n\u00061(z)\nfor the oscillaton from numerics (dotted lines) and the solution\nof Eqs. (21) (solid and dashed lines) at two di \u000berent instants.\nFigure 14(a) is for the left and Fig. 14(b) is for the right oscil-\nlaton.\nInterestingly, for \u001e1=0, increasing the extent of overlap\nor decreasing \u0001introduces a qualitatively new feature to the\ndynamics. Contrary to \u0001 =10l?case, an additional stationary\nferromagnetic soliton is formed, centerd at z=0, see Fig. 15.\nComparing Figs. 15 and 12, we see that decreasing \u0001reduces\nthe frequency of the internal oscillations and increases the ve-\nlocity of the final oscillatons. The ferromagnetic soliton ex-\nhibits a strong breathing character due to its dynamical forma-\ntion.\nFinally, we discuss the case of \u001e1=\u0019=2 rad. Unlike that of\n\u001e1=0, the two final oscillatons are symmetric [see Fig. 16].\nThere is no transient oscillatory dynamics at the initial stage of\nthe dynamics because of repulsion emerging from the initial\nphase di \u000berence of\u0019=2. Also, spin mixing happens indepen-\ndently in the left and right regions due to the initial repulsion.\nConsequently, we have independent and identical oscillatons\nmoving away from each other. Also, the noise does not a \u000bect10\nFigure 15. Dynamics of two overlapping polar bright solitons for\n\r=0:5,\u0001 = 5l?, ¯c0=~!?=\u00002, and\u001e1=0 rad. (a) show the\ndynamics ofj 1j2+j \u00001j2, (b) ofj 0j2and (c) of the total density\nn(z;t), respectively.\nFigure 16. Dynamics of two overlapping polar bright solitons for\n\r=0:5,\u0001 = 10l?, ¯c0=~!?=\u00002, and\u001e1=\u0019=2 rad. (a) show the\ndynamics ofj 1j2+j \u00001j2, (b) ofj 0j2and (c) of the total density\nn(z;t), respectively.\nthe dynamics in contrast to the case of \u001e1=0.IV . SUMMARY AND OUTLOOK\nIn summarizing, we have analyzed the dynamics of two\noverlapping polar bright solitons in spin-1 condensates, which\ndepends critically on the relative phase, the extent of overlap\nand the ratio between spin-independent and spin-dependent\ninteractions. The same dynamics of scalar solitons revealed\ninteresting scenarios, particularly the possibility of observing\natomic switching. Atomic switching can find applications in\nimplementing atom-based networks, identical to optical net-\nworks. Overlapping polar solitons resulted in non-trivial dy-\nnamics in spatial and spin degrees of freedom. For vanishing\nspin-dependent interactions, we observed Josephson like os-\ncillations of each component in an e \u000bective double-well po-\ntential created by the density of other components. For identi-\ncal spin-dependent and spin-independent interactions, we ob-\nserved the formation of four ferromagnetic solitons. In the\nlast case, when the spin-dependent interaction strength is half\nof the spin-independent one, the dynamics led to the forma-\ntion of two oscillatons. Strikingly, the properties of the final\nbright solitons can be easily tuned by the initial state and the\ninteraction parameters. Our studies o \u000ber a new possibility for\nengineering matter waves. Extending the above analysis to a\npair of overlapping ferromagnetic solitons and ferromagnetic-\npolar solitons may lead to exciting and completely new dy-\nnamics than those reported here. We also expect complex dy-\nnamics to emerge from overlapping more than two solitons.\nV . ACKNOWLEDGMENTS\nR.N. acknowledges support from DST-SERB for the Swar-\nnajayanti fellowship File No. SB /SJF/2020-21 /19. We ac-\nknowledge National Supercomputing Mission (NSM) for pro-\nviding computing resources of ’PARAM Brahma’ at IISER\nPune, which is implemented by C-DAC and supported by the\nMinistry of Electronics and Information Technology (MeitY)\nand Department of Science and Technology (DST), Govern-\nment of India. G. H. acknowledges the funding from DST\nIndia through an INSPIRE scholarship.\n[1] Y . Kawaguchi and M. Ueda, Phys. Rep. 520, 253 (2012).\n[2] M.-S. Chang, Q. Qin, W. Zhang, L. You, and M. S. Chapman,\nNature Physics 1, 111 (2005).\n[3] J. Kronj ¨ager, C. Becker, P. Navez, K. Bongs, and K. Sengstock,\nPhys. Rev. Lett. 97, 110404 (2006).\n[4] A. T. Black, E. Gomez, L. D. Turner, S. Jung, and P. D. Lett,\nPhys. Rev. Lett. 99, 070403 (2007).\n[5] Y . Liu, S. Jung, S. E. Maxwell, L. D. Turner, E. Tiesinga, and\nP. D. Lett, Phys. Rev. Lett. 102, 125301 (2009).\n[6] C. Klempt, O. Topic, G. Gebreyesus, M. Scherer, T. Henninger,\nP. Hyllus, W. Ertmer, L. Santos, and J. J. Arlt, Phys. Rev. Lett.\n103, 195302 (2009).\n[7] Y . Eto, H. Shibayama, H. Saito, and T. Hirano, Phys. Rev. A\n97, 021602 (2018).[8] Z. Chen, T. Tang, J. Austin, Z. Shaw, L. Zhao, and Y . Liu, Phys.\nRev. Lett. 123, 113002 (2019).\n[9] B. Evrard, A. Qu, J. Dalibard, and F. Gerbier, Phys. Rev. A\n103, L031302 (2021).\n[10] H. Pu, C. K. Law, S. Raghavan, J. H. Eberly, and N. P. Bigelow,\nPhys. Rev. A 60, 1463 (1999).\n[11] T. M. Bersano, V . Gokhroo, M. A. Khamehchi, J. D’Ambroise,\nD. J. Frantzeskakis, P. Engels, and P. G. Kevrekidis, Phys. Rev.\nLett. 120, 063202 (2018).\n[12] X. Chai, D. Lao, K. Fujimoto, R. Hamazaki, M. Ueda, and\nC. Raman, Phys. Rev. Lett. 125, 030402 (2020).\n[13] A. Farolfi, D. Trypogeorgos, C. Mordini, G. Lamporesi, and\nG. Ferrari, Phys. Rev. Lett. 125, 030401 (2020).\n[14] S. Lannig, C.-M. Schmied, M. Pr ¨ufer, P. Kunkel, R. Strohmaier,\nH. Strobel, T. Gasenzer, P. G. Kevrekidis, and M. K.11\nOberthaler, Phys. Rev. Lett. 125, 170401 (2020).\n[15] I.-K. Liu, S.-C. Gou, and H. Takeuchi, Phys. Rev. Research 2,\n033506 (2020).\n[16] X. Chai, D. Lao, K. Fujimoto, and C. Raman, Phys. Rev. Re-\nsearch 3, L012003 (2021).\n[17] L.-Z. Meng, Y .-H. Qin, and L.-C. Zhao, Communications\nin Nonlinear Science and Numerical Simulation 109, 106286\n(2022).\n[18] J. Ieda, T. Miyakawa, and M. Wadati, Phys. Rev. Lett. 93,\n194102 (2004).\n[19] J. Ieda, T. Miyakawa, and M. Wadati, J. Phys. Soc. Jpn. 73,\n2996 (2004).\n[20] L. Li, Z. Li, B. A. Malomed, D. Mihalache, and W. M. Liu,\nPhys. Rev. A 72, 033611 (2005).\n[21] P. Sza ´nkowski, M. Trippenbach, E. Infeld, and G. Rowlands,\nPhys. Rev. A 83, 013626 (2011).\n[22] M. Wang, B. Tian, W.-R. Shan, X. L ¨u, and Y .-S. Xue, Nonlin-\near Dyn. 69, 1137 (2012).\n[23] P. Szankowski, M. Trippenbach, E. Infeld, and G. Rowlands,\nPhys. Rev. Lett. 105, 125302 (2010).\n[24] W. Zhang, O. E. M ¨ustecaplıo ˘glu, and L. You, Phys. Rev. A 75,\n043601 (2007).[25] V . Saxena, I. Kourakis, G. Sanchez-Arriaga, and E. Siminos,\nPhys. Lett. A 377, 473 (2013).\n[26] V . V . Afanasjev and V . A. Vysloukh, J. Opt. Soc. Am. B 11,\n2385 (1994).\n[27] V . V . Afanasjev and N. Akhmediev, Phys. Rev. E 53, 6471\n(1996).\n[28] M. Shalaby and A. Barthelemy, Opt. Lett. 16, 1472 (1991).\n[29] M. Shalaby, F. Reynaud, and A. Barthelemy, Opt. Lett. 17, 778\n(1992).\n[30] J. S. Aitchison, A. M. Weiner, Y . Silberberg, D. E. Leaird, M. K.\nOliver, J. L. Jackel, and P. W. E. Smith, Opt. Lett. 16, 15 (1991).\n[31] S. G. I. and S. Mordechai, Science 286, 1518 (1999).\n[32] U. Al Khawaja, Phys. Rev. E 81, 056603 (2010).\n[33] J. H. V . Nguyen, P. Dyke, D. Luo, B. A. Malomed, and R. G.\nHulet, Nat. Phys. 10, 918 (2014).\n[34] T. P. Billam, S. L. Cornish, and S. A. Gardiner, Phys. Rev. A\n83, 041602 (2011).\n[35] L. M. Symes, R. I. McLachlan, and P. B. Blakie, Phys. Rev. E\n93, 053309 (2016).\n[36] K. L. Lee, N. B. Jørgensen, I.-K. Liu, L. Wacker, J. J. Arlt, and\nN. P. Proukakis, Phys. Rev. A 94, 013602 (2016)." }, { "title": "2109.06872v2.The_implications_of_high_BH_spins_on_the_origin_of_BH_BH_mergers.pdf", "content": "Draft version October 12, 2021\nPreprint typeset using L ATEX style emulateapj v. 12/16/11\nTHE IMPLICATIONS OF HIGH BH SPINS ON THE ORIGIN OF BH-BH MERGERS\nA. Olejak1, K. Belczynski1\n1Nicolaus Copernicus Astronomical Center, Polish Academy of Sciences, ul. Bartycka 18, 00-716 Warsaw, Poland,\n(aleksandra.olejak@wp.pl,chrisbelczynski@gmail.com)\nDraft version October 12, 2021\nABSTRACT\nThe LIGO/Virgo collaboration has reported 50 black hole|black hole (BH-BH) mergers and 8\ncandidates recovered from digging deeper into the detectors noise. The majority of these mergers\nhave low e\u000bective spins pointing toward low BH spins and e\u000ecient angular momentum transport\n(AM) in massive stars as proposed by several models (e.g., the Tayler-Spruit dynamo). However,\nout of these 58 mergers, 7 are consistent with having high e\u000bective spin parameter ( \u001fe\u000b>0:3).\nAdditionally, 2 events seem to have high e\u000bective spins sourced from the spin of the primary (more\nmassive) BH. These particular observations could be used to discriminate between the isolated binary\nand dynamical formation channels. It might seem that high BH spins point to a dynamical origin if\nAM in stars is e\u000ecient and forms low-spinning BHs. In such a case dynamical formation is required\nto produce second and third generations of BH-BH mergers with typically high-spinning BHs. Here\nwe show, however, that isolated binary BH-BH formation naturally reproduces such highly spinning\nBHs. Our models start with e\u000ecient AM in massive stars that is needed to reproduce the majority\nof BH-BH mergers with low e\u000bective spins. Later, some of the binaries are subject to a tidal spin-up\nallowing the formation of a moderate fraction ( \u001810%) of BH-BH mergers with high e\u000bective spins\n(\u001fe\u000b&0:4\u00000:5). In addition, isolated{binary evolution can produce a small fraction of BH-BH\nmergers with almost maximally spinning primary BHs. Therefore, the formation scenario of these\natypical BH-BH mergers remains to be found.\nSubject headings: stars: black holes, compact objects, massive stars\n1.INTRODUCTION\nThe LIGO/Virgo collaboration has announced detec-\ntion of gravitational waves from \u001850 double black\nhole (BH-BH) mergers (Abbott et al. 2019a,b; Fish-\nbach & Holz 2020; Abbott et al. 2021a). Additional\n8 BH-BH merger candidates have been recently re-\nported (Abbott et al. 2021b). The majority of all\nthese events have low e\u000bective spins parameters: \u001fe\u000b=\nm1a1cos\u00121+m2a2cos\u00122\nm1+m2\u00190;wheremiare BH masses,\nai=cJi=Gm2\niare dimensionless BH spin magnitudes\n(Jibeing the BH angular momentum (AM), cthe speed\nof light,Gthe gravitational constant), and \u0012iare angles\nbetween the individual BH spins and the system orbital\nAM.\nHowever, among the dectections there are also several\nBH-BH mergers which are characterized by higher (non-\nzero) positive e\u000bective spins. In Table 1 we list the pa-\nrameters of the \fve BH-BH mergers with highest e\u000bective\nspins reported by Abbott et al. (2021a) with additional\ntwo high e\u000bective spin systems reported by Abbott et al.\n(2021b).\nThe formation of close BH-BH systems is an open issue\nwith several formation channels proposed and discussed\nin the context of the LIGO/Virgo mergers. The major\nformation scenarios include the isolated binary evolution\n(Bond & Carr 1984; Tutukov & Yungelson 1993; Lipunov\net al. 1997; Voss & Tauris 2003; Belczynski et al. 2010b;\nDominik et al. 2012; Kinugawa et al. 2014; Hartwig et al.\n2016; de Mink & Mandel 2016; Mandel & de Mink 2016;\nMarchant et al. 2016; Spera et al. 2016; Belczynski et al.\n2016a; Eldridge & Stanway 2016; Woosley 2016; van den\nHeuvel et al. 2017; Stevenson et al. 2017; Kruckow et al.TABLE 1\nBH-BH mergers with high effective spins\nNo. Namea\u001fe\u000bm1m2a1\n1 GW190517 0 :52+0:19\n\u00000:1937:4+11:7\n\u00007:625:3+7:0\n\u00007:3{\n2 GW170729 0 :37+0:21\n\u00000:2550:2+16:2\n\u000010:234:0+9:1\n\u000010:1{\n3 GW190620 0 :33+0:22\n\u00000:2557:1+16:0\n\u000012:735:5+12:2\n\u000012:3{\n4 GW190519 0 :31+0:20\n\u00000:2266:0+10:7\n\u000012:040:5+11:0\n\u000011:1{\n5 GW190706 0 :28+0:26\n\u00000:2967:0+14:6\n\u000013:338:2+14:6\n\u000013:3{\n6 GW190403 0 :70+0:15\n\u00000:2788:0+28:2\n\u000032:922:1+23:8\n\u00009:00:92+0:07\n\u00000:22\n7 GW190805 0 :35+0:30\n\u00000:3648:2+17:5\n\u000012:532:0+13:4\n\u000011:40:74+0:22\n\u00000:60\na: Names are abbreviated. We include candidate detections\nas full astrophysical events. Parameters of \frst 5 events are\nfrom original LIGO/Virgo analysis (Abbott et al. 2021a), while\nthe remaining 2 are from deeper search into the detectors noise\n(Abbott et al. 2021b).\n2018; Hainich et al. 2018; Marchant et al. 2018; Spera\net al. 2019; Neijssel et al. 2019; du Buisson et al. 2020;\nBavera et al. 2020, 2021; Qin et al. 2021), the dense stel-\nlar system dynamical channel (Portegies Zwart & McMil-\nlan 2000; Miller & Hamilton 2002b,a; Portegies Zwart\net al. 2004; G ultekin et al. 2004, 2006; O'Leary et al.\n2007; Sadowski et al. 2008; Downing et al. 2010; An-\ntonini & Perets 2012a; Benacquista & Downing 2013;\nMennekens & Vanbeveren 2014; Bae et al. 2014; Chat-\nterjee et al. 2016; Mapelli 2016; Hurley et al. 2016; Ro-\ndriguez et al. 2016; VanLandingham et al. 2016; Askar\net al. 2017; Arca-Sedda & Capuzzo-Dolcetta 2017; Sam-\nsing 2017; Morawski et al. 2018; Banerjee 2018; Di Carlo\net al. 2019; Zevin et al. 2019; Rodriguez et al. 2018;\nPerna et al. 2019; Kremer et al. 2020), isolated multi-arXiv:2109.06872v2 [astro-ph.HE] 11 Oct 20212\nple (triple, quadruple) systems (Antonini et al. 2017;\nSilsbee & Tremaine 2017; Arca-Sedda et al. 2018; Liu &\nLai 2018; Fragione & Kocsis 2019), mergers of binaries in\ngalactic nuclei (Antonini & Perets 2012b; Hamers et al.\n2018; Hoang et al. 2018; Fragione et al. 2019) and pri-\nmordial BH formation (Sasaki et al. 2016; Green 2017;\nClesse & Garc\u0013 \u0010a-Bellido 2017; Carr & Silk 2018; De Luca\net al. 2020).\nBH spins and their orientations can play an important\nrole in distinguishing between various BH-BH formation\nmodels. If the BH spins are not small, then their orienta-\ntion may possibly distinguish between a binary evolution\norigin (predominantly aligned spins) and dynamical for-\nmation channels (more or less isotropic distribution of\nspin orientations). If the BHs formed out of stars have\nsmall spins (Spruit 2002; Zaldarriaga et al. 2017; Ho-\ntokezaka & Piran 2017; Fuller et al. 2019; Qin et al. 2019;\nOlejak et al. 2020; Bavera et al. 2020; Belczynski et al.\n2020) then BH-BH mergers with high e\u000bective spins may\nchallenge their isolated evolution origin. In dense stellar\nclusters, BHs may merge several times easily producing\nBHs with high spins and making a dynamical channel a\nprime site for such events (Gerosa & Berti 2017; Fishbach\net al. 2017). However, the assumption about the BH na-\ntal spin (and the AM transport e\u000eciency) also plays a\nrole in the e\u000bective spin distribution for the dynamical\nchannel (Banerjee 2021).\nIn this study we show that the current understanding\nof stellar/binary astrophysics (Belczynski et al. 2021) and\nthe degeneracy between the di\u000berent formation channels\ndo not allow for such a simple test of the origin of the\nLIGO/Virgo BH-BH mergers. To demonstrate this we\nshow that although the isolated binary evolution channel\nproduces mostly BH-BH mergers with low e\u000bective spins,\na small but signi\fcant fraction of mergers is expected to\nhave moderate or even high e\u000bective spins. Despite the\nassumption that stars slow down their rotation due to\ne\u000ecient AM transport, we \fnd that tidal interactions\nare capable of spinning up some stars allowing formation\nof rapidly spinning BHs (Detmers et al. 2008; Kushnir\net al. 2017; Qin et al. 2018).\n2.METHOD\nWe use the population synthesis code StarTrack (Bel-\nczynski et al. 2002, 2008) with a model of star forma-\ntion rates and metallicity distribution based on Madau\n& Dickinson (2014) described in Belczynski et al. (2020).\nWe employ the delayed core-collapse supernova (SN) en-\ngine for neutron star/BH mass calculation (Fryer et al.\n2012), with weak mass loss from pulsation pair insta-\nbility supernovae (Belczynski et al. 2016b). BH na-\ntal kicks are calculated from a Maxwellian distribution\nwith\u001b= 265 km s\u00001and decreased by fallback during\ncore-collapse; this makes a signi\fcant fraction of BHs\nform without a natal kick (Mirabel & Rodrigues 2003).\nWe assume our standard wind losses for massive O/B\nstars (Vink et al. 2001) and LBV winds (speci\fc pre-\nscriptions for these winds are listed in Sec. 2.2 of Bel-\nczynski et al. 2010a). BH natal spins are calculated under\nthe assumption that AM in massive stars is transported\nby the Tayler-Spruit magnetic dynamo (Spruit 2002) as\nadopted in the MESA stellar evolutionary code (Paxton\net al. 2015). Such BH natal spins take values in the range\na20:05\u00000:15 (see Belczynski et al. 2020). Note thatthe modi\fed classic Tayler-Spruit dynamo with a di\u000ber-\nent non-linear saturation mechanism of the Tayler insta-\nbility (Fuller et al. 2019; Fuller & Ma 2019) causes larger\nmagnetic \feld amplitudes, more e\u000ecient AM transport\nand even lower \fnal natal spins ( a\u00180:01). BH spin\nmay be increased if the immediate BH progenitors (Wolf-\nRayet: WR) stars in close binaries are subject to tidal\nspin-up. In our calculations for BH-WR, WR-BH and\nWR-WR binary systems with orbital periods in the range\nPorb= 0:1\u00001:3 d the BH natal spin magnitude is \ft from\nWR star spun-up MESA models (see eq.15 of Belczyn-\nski et al. (2020)), while for systems with Porb<0:1d\nthe BH spin is taken to be equal to a= 1. BH spins\nmay also be increased by accretion in binary systems.\nWe treat accretion onto a compact object during Roche\nlobe over\row (RLOF) and from stellar winds using the\nanalytic approximations presented in King et al. (2001)\nand Mondal et al. (2020). In the adopted approach the\naccumulation of matter on a BH is very ine\u000ecient so\naccretion does not noticeably a\u000bect the \fnal BH spin.\nHowever note that, e.g. van Son et al. (2020) or Bavera\net al. (2021) tested di\u000berent super Eddington accretion\nprescriptions \fnding that some BHs may be signi\fcantly\nspun-up by accretion.\nFor common the envelope (CE) evolution we assume\na 100% (\u000bCE= 1) orbital energy transfer for CE ejec-\ntion and we adopt 5% Bondi accretion rate onto the BHs\nduring CE (Ricker & Taam 2008; MacLeod & Ramirez-\nRuiz 2015; MacLeod et al. 2017). During the stable\nRLOF (whether it is a thermal- or nuclear-timescale mass\ntransfer: TTMT/NTMT) we adopt the following input\nphysics. If an accretor is a compact object (neutron star\nor BH) we allow for super-Eddington accretion with ex-\ncess transferred mass lost with an AM speci\fc to the\naccretor (Mondal et al. 2020). In all other cases, we al-\nlow a fraction of the transferred mass of fa= 0:5 to be\nlost from the binary with a speci\fc AM of the binary\norbitjloss= 1:0 (expressed in units of 2 \u0019A2=Porb,Abe-\ning an orbital separation; see eq. 33 of Belczynski et al.\n(2008)).\nRLOF stability is an important issue in the context of\nBH-BH system formation in the framework of the iso-\nlated binary evolution (Neijssel et al. 2019; Olejak et al.\n2021; Gallegos-Garcia et al. 2021; Belczynski et al. 2021).\nIn the standard StarTrack evolution we impose rather\nliberal limits for CE (dynamical-timescale RLOF) to de-\nvelop (see Belczynski et al. (2008): binaries with a donor\nstar more massive than 2 \u00003 times the mass of the ac-\ncretor are subject to CE. In this model (for simplicity\ntagged here as CE model) the vast majority of BH-BH\nmergers form through CE evolution, although we \fnd\nsome cases ( .1%) of BH-BH merger formation without\nany CE event. In the alternative model (non-CE model,\ndetailed description in Olejak et al. 2021) we allow CE\nto be suppressed for some systems even with mass ra-\ntio as high as 6 \u00008 (Pavlovskii et al. 2017). In this\nmodel the majority of the BH-BH mergers form without\nany CE event (the orbital decrease is obtained through\nangular momentum loss during stable RLOF), although\nsome (<10%) BH-BH mergers form with the assistance\nof CE.\nFor each model we calculate the evolution of 64 mil-\nlion massive, Population I/II binary systems. We use\nthe star formation history and chemical evolution of the3\nUniverse to obtain the BH-BH merger properties within\nan approximate reach of LIGO/Virgo (redshift z < 1).\nWe use the same method as described in Belczynski et al.\n(2020).\n3.RESULTS\nFigure 1 shows a typical example of binary system evo-\nlution without a CE phase leading to the formation a\nBH-BH merger with a tidally spun-up primary BH (re-\nstricted RLOF stability criteria; Olejak et al. 2021). The\nrather unequal-mass massive stellar system (112 M\fand\n68M\f) with a metallicity of Z= 0:002 goes through two\nRLOF events. The RLOF I is initiated by the more mas-\nsive star; \frst by an NTMT when the donor is still on\nthe main-sequence and then through a TTMT when the\ndonor evolves o\u000b main-sequence. After the RLOF I, the\nsystem mass ratio is reversed: the initially more mas-\nsive star lost over 80% of its mass while the companion\ngained\u001840M\f. Next, the initially more massive star\nends its evolution directly collapsing to the less{massive\n(secondary) BH with a mass of m2= 15M\fand spin\na2= 0:14. When the companion star expands, it ini-\ntiates a second stable RLOF. At the onset of RLOF II\nthe system has highly unequal masses: the donor is al-\nmost 6:5 times more massive than the BH. The thermal\ntimescale for a donor with mass Mdon\u001997M\f, radius\nRdon\u0019300R\fand luminosity Ldon\u00193\u0002106L\f(pa-\nrameters at the RLOF II onset1) calculated with the\nformula by Kalogera & Webbink (1996), is \u001cth\u0019330\nyr:It corresponds to a very high mass transfer rate\n_M=Mdon=\u001cth\u00190:3 M\fyr\u00001which does not allow the\nBH to accrete much mass (despite the fact that we allow\nfor super-Eddington accretion). Half of the donors mass\nis lost from the binary with the speci\fc AM of the BH (as\nthe matter was transferred to the vicinity of the BH ac-\ncretor). This has a huge e\u000bect on the orbital separation\nwhich decreases from A= 467R\fto onlyA= 7:1R\f.\nAfter RLOF II the binary consists of a BH and a WR star\nthat are close enough to allow for the tidal spin-up of the\nWR star. Finally, the WR star directly collapses to the\nmore massive (primary) BH with a mass m1= 36M\f\nand spina1= 0:68. The BH-BH system merges after\n\u001867 Myr.\nFigure 2 shows a typical CE evolution scenario (stan-\ndard StarTrack RLOF stability criteria) leading to the\nformation a BH-BH merger with both BHs spun-up by\ntidal interactions. At the beginning, the binary system\nof two\u001836M\fstars with Z= 0:0025 is on a wide\n(A\u00191340R\f) and eccentric orbit ( e= 0:1). When\nthe initially more massive star expands the system goes\nthrough a stable RLOF, after which the donor looses\nits H-rich envelope and the orbit circularizes. Soon af-\nter RLOF I, the system goes through another (unstable)\nRLOF initiated by the initially less massive companion\nstar. The ensuing CE evolution leads to signi\fcant or-\nbital contraction from A= 3100R\ftoA= 4:5R\fand\nleaves two WR stars subject to strong tidal interactions.\nBoth stars end their evolution at a similar time form-\ning via supernovae explosions two \u00189M\fBHs. At\nthe formation, both BHs get signi\fcant natal kicks that\n1Such parameter values are inline with other predictions for\nmassive stars e.g. using Geneva stellar evolution code (Yusof et al.\n2013).\nFig. 1.| Typical example of non-CE evolutionary scenario lead-\ning to the formation of BH-BH merger with tidally spun-up pri-\nmary:a1= 0:68 and\u001fe\u000b= 0:52. Binary system goes through two\nphases of RLOF with episodes of nuclear and thermal timescale\nmass transfer. RLOF I ends with the system mass ratio reversal.\nAfter RLOF II the system orbital separation signi\fcantly decreases\nand WR star is a subject to tidal spin-up by a BH. Soon there-\nafter the close BH-BH system is formed with a short merger time\nof\u001867 Myr (see Sec. 3).\nmakes the system orbit larger A\u001919R\fand eccentric\ne= 0:44, leading to a merger time of \u00186:7 Gyr.\nIn Table 2 we present the statistical spin proper-\nties of BH-BH systems merging at redshifts z < 1\nfor the two tested RLOF stability criteria models. In\nthe rows 1\u00006 we list the percentage of the BH-BH\nmergers with e\u000bective{spin parameter values \u001fe\u000b>\n0:0;0:1;0:2;0:3;0:4;0:5. In the rows 7\u00009 we list the\npercentages of BH-BH mergers with a highly spinning\nprimary BH a1>0:5;0:7;0:9 while the rows 10 \u000012\ngive the percentages of mergers with a highly spinning\nsecondary BH a2>0:5;0:7;0:9. The full distribution of\nthe primary{spin, the secondary{spin and the e\u000bective{\nspin parameter for both the CE and non-CE evolution,\nis plotted in Figure 3 in APPENDIX A.\n4.DISCUSSION AND CONCLUSIONS\nThe rapidly increasing number of detected BH-BH\nmergers does allow for some general population state-\nments (Roulet et al. 2021; Galaudage et al. 2021; Abbott\net al. 2021b). It appears that (i)majority (\u001870\u000090%) of\nBH-BH mergers have low e\u000bective spins consistent with\n\u001fe\u000b\u00190 and that (ii)small fraction (\u001810\u000030%) of\nmergers have positive non-zero spins that can be as high\nas\u001fe\u000b&0:5. Additionally, the population is consistent\nwith (iii)no systems having negative e\u000bective spins and4\nFig. 2.| Typical example of evolutionary scenario with CE\nphase leading to the formation of BH-BH merger with a1= 0:79,\na2= 0:79 and\u001fe\u000b= 0:77. First, the binary system goes through\nstable RLOF phase with episodes of nuclear and thermal timescale\nmass transfer initiated by the initially more massive star. Then\ninitially less massive star expands and initiates CE, after which\nthe orbital separation is signi\fcantly decreased. After CE, binary\nhosts two compact WR stars that are subject to tidal spin-up.\nBoth stars explode as supernovae and form BHs on eccentric orbit\nwith merger time of \u00186:7 Gyr (see Sec. 3).\nTABLE 2\nPredictions for BH-BH mergers from binary evolution\nNo. conditionaCE model non-CE model\n1\u001fe\u000b>0:0 97% 93%\n2\u001fe\u000b>0:1 95% 85%\n3\u001fe\u000b>0:2 70% 60%\n4\u001fe\u000b>0:3 36% 39%\n5\u001fe\u000b>0:4 10% 21%\n6\u001fe\u000b>0:5 2% 7%\n7a1>0:5 3% 34%\n8a1>0:7 2% 15%\n9a1>0:9 1% 1%\n10a2>0:5 52% 11%\n11a2>0:7 33% 7%\n12a2>0:9 12% 2%\na: We list fractions of BH-BH mergers (redshift z<1) produced in\nour two population synthesis models satisfying a given condition.(iv)a not isotropic distribution of e\u000bective spins (which\ncould indicate dynamical origin). Finally, (v)there is\nat least one case of a primary BH (more massive) in a\nBH-BH merger with very high spin ( a1>0:7 at 90%\ncredibility). These properties are noted to be broadly\nconsistent with BH-BH mergers being formed in an iso-\nlated binary evolution.\nIn our study we have tested whether we can reproduce\nthe above spin characteristics with our binary evolution\nmodels that employ e\u000ecient AM transport in massive\nstars and that impose tidal spin-up of compact massive\nWolf-Rayet stars in close binaries. The two presented\nmodels employ our standard input physics but allow for\nthe formation of BH-BH mergers assisted either by a CE\nor by a stable RLOF. We \fnd that the observed popula-\ntion and its spin characteristics ( i{v) is consistent with\nour isolated{binary{evolution predictions (see Tab. 2).\nIn particular, we \fnd that the majority of BH-BH merg-\ners have small positive e\u000bective spins: \u001870% mergers\nhave 0< \u001f e\u000b<0:3 (e\u000ecient AM transport), while a\nsmall fraction have signi\fcant spins: 36 \u000039% mergers\nhave\u001fe\u000b>0:3 and 2\u00007% mergers have \u001fe\u000b>0:5 (tidal\nspin-up). The fraction of systems with negative e\u000bective\nspins is small (3\u00007%) as most BHs do not receive strong\nnatal kicks in our simulations. Individual BH spins can\nreach high values. A large fraction (11 \u000052%) of sec-\nondary BHs may have signi\fcant spin values ( a2>0:5)\nas it is the less massive stars that are most often sub-\nject to tidal spin-up. Nevertheless, primary BHs may\nalso form with high spins (3 \u000034% witha1>0:5) if\nboth stars have similar masses and both are subject to\ntidal spin-up (see Fig. 2) or due to mass ratio reversal\ncaused by the RLOF (see Fig. 1). We also note the for-\nmation of a small fraction of almost maximally spinning\nBHs: 2\u000012% fora2>0:9 (secondary BH) and 1% for\na1>0:9 (primary BH). These results on e\u000bective spins\nand individual BH spins are consistent with the current\nLIGO/Virgo population of BH-BH mergers. Note that\nQin et al. (2021) came to di\u000berent conclusions, \fnding\nthe high-spinning detections challenging for the Tayler-\nSpruit dynamo, especially for the unequal mass event\nwith a high spinning primary (GW190403). Our non-CE\nmodel reproduces this type of mergers due to the mass\nratio reversal (see Fig. 1). In this channel, at the onset\nof the second stable RLOF, the donor may be even 5-6\ntimes more massive than the accretor, ending as an un-\nequal mass ( q\u00140:4) BH-BH merger. Qin et al. (2021)\nhave not considered the case of a stable RLOF in such\nunequal mass systems.\nThe above fractions correspond to just two di\u000berent\nmodes of spinning-up during the classical isolated binary\nBH-BH formation. Had we varied several other factors\nthat in\ruence BH spins and their orientations in BH-\nBH mergers, the ranges of these fractions would have\nbroadened. Some obvious physical processes that can af-\nfect BH spins and their orientations include: initial star\nspin alignment (or lack thereof) with the binary AM, the\nalignment of stellar spins (or lack thereof) during RLOF\nphases, the treatment of accretion, the initial mass ra-\ntio distribution that can alter the ratio of systems going\nthrough stable and unstable (CE) RLOF, and the natal\nkicks that can misalign spin orientations. Above all, the\nthree major uncertainties include the initial stellar rota-\ntion of stars forming BHs, the e\u000eciency of AM transport5\nand the strength of tides in close binary systems. All of\nthe above are only weakly constrained. Note that this\nis a proof-of-principle study that is limited only to BH\nspins in BH-BH mergers. In particular, we did not try\nto match BH masses and BH-BH merger rates for the\nhighly spinning LIGO/Virgo BHs. In this study we have\nonly shown that it is possible to produce highly spinning\nBHs by tidal interactions of stars in close binaries in evo-\nlution that includes and does not include CE. Our two\nexamples of evolution (Fig. 1 and 2) have much smaller\nmasses than the LIGO/Virgo mergers with highly spin-\nning BHs (Tab. 1). Note, however, that we have not\nused here the input physics that allows for the formation\nof BHs with mass over 50 M\f. Such model is already\nincorporated and tested within our population synthe-\nsis code (Belczynski 2020). An attempt to match allobserved parameters simultaneously is projected to hap-\npen in the future when LIGO/Virgo will deliver a larger\nsample of highly spinning BHs.\nGiven the results presented in this study, alas limited\nonly to BH spins, we conclude that (i)the isolated binary\nevolution channel reproduces well the BH spins of the\nLIGO/Virgo mergers (ii)if, in fact, the binary channel\nis producing the majority of the LIGO/Virgo BH-BH\nmergers, then this indicates that the AM transport is\ne\u000ecient in massive stars and the tidal interactions in\nclose binaries are strong.\nWe thank the anonymous reviewer, Jean-Pierre La-\nsota, Ilya Mandel and Sambaran Banerjee for their useful\ncomments on the manuscript. KB and AO acknowledge\nsupport from the Polish National Science Center (NCN)\ngrant Maestro (2018/30/A/ST9/00050).\nREFERENCES\nAbbott, B. P., Abbott, R., Abbott, T. D., Abraham, S., LIGO\nScienti\fc Collaboration, & Virgo Collaboration. 2019a, ApJ,\n882, L24\n|. 2019b, Physical Review X, 9, 031040\nAbbott, R., et al. 2021a, Physical Review X, 11, 021053\n|. 2021b, arXiv e-prints, arXiv:2108.01045\nAntonini, F., & Perets, H. B. 2012a, ApJ, 757, 27\n|. 2012b, ApJ, 757, 27\nAntonini, F., Toonen, S., & Hamers, A. S. 2017, ApJ, 841, 77\nArca-Sedda, M., & Capuzzo-Dolcetta, R. 2017, ArXiv e-prints\nArca-Sedda, M., Li, G., & Kocsis, B. 2018, arXiv e-prints,\narXiv:1805.06458\nAskar, A., Szkudlarek, M., Gondek-Rosi\u0013 nska, D., Giersz, M., &\nBulik, T. 2017, MNRAS, 464, L36\nBae, Y.-B., Kim, C., & Lee, H. M. 2014, MNRAS, 440, 2714\nBanerjee, S. 2018, MNRAS, 473, 909\n|. 2021, MNRAS, 500, 3002\nBavera, S. S., et al. 2020, A&A, 635, A97\n|. 2021, A&A, 647, A153\nBelczynski, K. 2020, ApJ, 905, L15\nBelczynski, K., Bulik, T., Fryer, C. L., Ruiter, A., Valsecchi, F.,\nVink, J. S., & Hurley, J. R. 2010a, ApJ, 714, 1217\nBelczynski, K., Dominik, M., Bulik, T., O'Shaughnessy, R., Fryer,\nC. L., & Holz, D. E. 2010b, ApJ, 715, L138\nBelczynski, K., Holz, D. E., Bulik, T., & O'Shaughnessy, R.\n2016a, Nature, 534, 512\nBelczynski, K., Kalogera, V., & Bulik, T. 2002, ApJ, 572, 407\nBelczynski, K., Kalogera, V., Rasio, F. A., Taam, R. E., Zezas,\nA., Bulik, T., Maccarone, T. J., & Ivanova, N. 2008, ApJS, 174,\n223\nBelczynski, K., et al. 2016b, A&A, 594, A97\n|. 2020, A&A, 636, A104\n|. 2021, arXiv e-prints, arXiv:2108.10885\nBenacquista, M. J., & Downing, J. M. B. 2013, Living Reviews in\nRelativity, 16, 4\nBond, J. R., & Carr, B. J. 1984, MNRAS, 207, 585\nCarr, B., & Silk, J. 2018, MNRAS, 478, 3756\nChatterjee, S., Rodriguez, C. L., Kalogera, V., & Rasio, F. A.\n2016, ArXiv e-prints\nClesse, S., & Garc\u0013 \u0010a-Bellido, J. 2017, Physics of the Dark\nUniverse, 15, 142\nDe Luca, V., Desjacques, V., Franciolini, G., Pani, P., & Riotto,\nA. 2020, arXiv e-prints, arXiv:2009.01728\nde Mink, S. E., & Mandel, I. 2016, MNRAS, 460, 3545\nDetmers, R. G., Langer, N., Podsiadlowski, P., & Izzard, R. G.\n2008, A&A, 484, 831\nDi Carlo, U. N., Giacobbo, N., Mapelli, M., Pasquato, M., Spera,\nM., Wang, L., & Haardt, F. 2019, MNRAS, 487, 2947\nDominik, M., Belczynski, K., Fryer, C., Holz, D., Berti, B., Bulik,\nT., Mandel, I., & O'Shaughnessy, R. 2012, ApJ, 759, 52\nDowning, J. M. B., Benacquista, M. J., Giersz, M., & Spurzem,\nR. 2010, MNRAS, 407, 1946\ndu Buisson, L., et al. 2020, arXiv e-prints, arXiv:2002.11630\nEldridge, J. J., & Stanway, E. R. 2016, MNRAS, 462, 3302\nFishbach, M., & Holz, D. E. 2020, ApJ, 891, L27\nFishbach, M., Holz, D. E., & Farr, B. 2017, ApJ, 840, L24\nFragione, G., Grishin, E., Leigh, N. W. C., Perets, H. B., &\nPerna, R. 2019, MNRAS, 488, 47\nFragione, G., & Kocsis, B. 2019, MNRAS, 486, 4781Fryer, C. L., Belczynski, K., Wiktorowicz, G., Dominik, M.,\nKalogera, V., & Holz, D. E. 2012, ApJ, 749, 91\nFuller, J., & Ma, L. 2019, ApJ, 881, L1\nFuller, J., Piro, A. L., & Jermyn, A. S. 2019, MNRAS\nGalaudage, S., Talbot, C., Nagar, T., Jain, D., Thrane, E., &\nMandel, I. 2021, arXiv e-prints, arXiv:2109.02424\nGallegos-Garcia, M., Berry, C. P. L., Marchant, P., & Kalogera,\nV. 2021, arXiv e-prints, arXiv:2107.05702\nGerosa, D., & Berti, E. 2017, Phys. Rev. D, 95, 124046\nGreen, A. M. 2017, ArXiv e-prints\nG ultekin, K., Miller, M. C., & Hamilton, D. P. 2004, ApJ, 616,\n221\n|. 2006, ApJ, 640, 156\nHainich, R., et al. 2018, A&A, 609, A94\nHamers, A. S., Bar-Or, B., Petrovich, C., & Antonini, F. 2018,\nApJ, 865, 2\nHartwig, T., Volonteri, M., Bromm, V., Klessen, R. S., Barausse,\nE., Magg, M., & Stacy, A. 2016, MNRAS, 460, L74\nHoang, B.-M., Naoz, S., Kocsis, B., Rasio, F. A., & Dosopoulou,\nF. 2018, ApJ, 856, 140\nHotokezaka, K., & Piran, T. 2017, ArXiv e-prints\nHurley, J. R., Sippel, A. C., Tout, C. A., & Aarseth, S. J. 2016,\nMNRAS, 33, e036\nKalogera, V., & Webbink, R. F. 1996, Astrophys. J., 458, 301\nKing, A. R., Davies, M. B., Ward, M. J., Fabbiano, G., & Elvis,\nM. 2001, ApJ, 552, L109\nKinugawa, T., Inayoshi, K., Hotokezaka, K., Nakauchi, D., &\nNakamura, T. 2014, MNRAS, 442, 2963\nKremer, K., et al. 2020, ApJS, 247, 48\nKruckow, M. U., Tauris, T. M., Langer, N., Kramer, M., &\nIzzard, R. G. 2018, ArXiv e-prints\nKushnir, D., Zaldarriaga, M., Kollmeier, J. A., & Waldman, R.\n2017, MNRAS, 467, 2146\nLipunov, V. M., Postnov, K. A., & Prokhorov, M. E. 1997,\nAstronomy Letters, 23, 492\nLiu, B., & Lai, D. 2018, ApJ, 863, 68\nMacLeod, M., Antoni, A., Murguia-Berthier, A., Macias, P., &\nRamirez-Ruiz, E. 2017, ApJ, 838, 56\nMacLeod, M., & Ramirez-Ruiz, E. 2015, ApJ, 803, 41\nMadau, P., & Dickinson, M. 2014, ARA&A, 52, 415\nMandel, I., & de Mink, S. E. 2016, MNRAS, 458, 2634\nMapelli, M. 2016, MNRAS, 459, 3432\nMarchant, P., Langer, N., Podsiadlowski, P., Tauris, T. M., &\nMoriya, T. J. 2016, A&A, 588, A50\nMarchant, P., Renzo, M., Farmer, R., Pappas, K. M. W., Taam,\nR. E., de Mink, S., & Kalogera, V. 2018, arXiv e-prints\nMennekens, N., & Vanbeveren, D. 2014, A&A, 564, A134\nMiller, M. C., & Hamilton, D. P. 2002a, ApJ, 576, 894\n|. 2002b, MNRAS, 330, 232\nMirabel, I. F., & Rodrigues, I. 2003, Science, 300, 1119\nMondal, S., Belczy\u0013 nski, K., Wiktorowicz, G., Lasota, J.-P., &\nKing, A. R. 2020, MNRAS, 491, 2747\nMorawski, J., Giersz, M., Askar, A., & Belczynski, K. 2018,\nArXiv e-prints\nNeijssel, C. J., et al. 2019, MNRAS, 490, 3740\nO'Leary, R. M., O'Shaughnessy, R., & Rasio, F. A. 2007,\nPhys. Rev. D, 76, 061504\nOlejak, A., Belczynski, K., & Ivanova, N. 2021, A&A, 651, A100\nOlejak, A., Fishbach, M., Belczynski, K., Holz, D. E., Lasota,\nJ. P., Miller, M. C., & Bulik, T. 2020, ApJ, 901, L396\nPavlovskii, K., Ivanova, N., Belczynski, K., & Van, K. X. 2017,\nMNRAS, 465, 2092\nPaxton, B., et al. 2015, ApJS, 220, 15\nPerna, R., Wang, Y.-H., Farr, W. M., Leigh, N., & Cantiello, M.\n2019, ApJ, 878, L1\nPortegies Zwart, S. F., Baumgardt, H., Hut, P., Makino, J., &\nMcMillan, S. L. W. 2004, Nature, 428, 724\nPortegies Zwart, S. F., & McMillan, S. L. W. 2000, ApJ, 528, L17\nQin, Y., Fragos, T., Meynet, G., Andrews, J., S\u001crensen, M., &\nSong, H. F. 2018, A&A, 616, A28\nQin, Y., Marchant, P., Fragos, T., Meynet, G., & Kalogera, V.\n2019, ApJ, 870, L18\nQin, Y., Wang, Y.-Z., Dong-Hong, Wu, Meynet, G., & Song, H.\n2021, arXiv e-prints, arXiv:2108.04821\nRicker, P. M., & Taam, R. E. 2008, ApJ, 672, L41\nRodriguez, C. L., Amaro-Seoane, P., Chatterjee, S., Kremer, K.,\nRasio, F. A., Samsing, J., Ye, C. S., & Zevin, M. 2018,\nPhys. Rev. D, 98, 123005\nRodriguez, C. L., Haster, C.-J., Chatterjee, S., Kalogera, V., &\nRasio, F. A. 2016, ApJ, 824, L8\nRoulet, J., Chia, H. S., Olsen, S., Dai, L., Venumadhav, T.,\nZackay, B., & Zaldarriaga, M. 2021, arXiv e-prints,\narXiv:2105.10580\nSadowski, A., Belczynski, K., Bulik, T., Ivanova, N., Rasio, F. A.,\n& O'Shaughnessy, R. 2008, ApJ, 676, 1162\nSamsing, J. 2017, ArXiv e-printsSasaki, M., Suyama, T., Tanaka, T., & Yokoyama, S. 2016, ArXiv\ne-prints\nSilsbee, K., & Tremaine, S. 2017, ApJ, 836, 39\nSpera, M., Giacobbo, N., & Mapelli, M. 2016,\nMem. Soc. Astron. Italiana, 87, 575\nSpera, M., Mapelli, M., Giacobbo, N., Trani, A. A., Bressan, A.,\n& Costa, G. 2019, MNRAS, 485, 889\nSpruit, H. C. 2002, A&A, 381, 923\nStevenson, S., Vigna-G\u0013 omez, A., Mandel, I., Barrett, J. W.,\nNeijssel, C. J., Perkins, D., & de Mink, S. E. 2017, Nature\nCommunications, 8, 14906\nTutukov, A. V., & Yungelson, L. R. 1993, MNRAS, 260, 675\nvan den Heuvel, E. P. J., Portegies Zwart, S. F., & de Mink, S. E.\n2017, MNRAS, 471, 4256\nvan Son, L. A. C., et al. 2020, ApJ, 897, 100\nVanLandingham, J. H., Miller, M. C., Hamilton, D. P., &\nRichardson, D. C. 2016, ApJ, 828, 77\nVink, J. S., de Koter, A., & Lamers, H. J. G. L. M. 2001, A&A,\n369, 574\nVoss, R., & Tauris, T. M. 2003, MNRAS, 342, 1169\nWoosley, S. E. 2016, ApJ, 824, L10\nYusof, N., et al. 2013, MNRAS, 433, 1114\nZaldarriaga, M., Kushnir, D., & Kollmeier, J. A. 2017, ArXiv\ne-prints\nZevin, M., Samsing, J., Rodriguez, C., Haster, C.-J., &\nRamirez-Ruiz, E. 2019, ApJ, 871, 91\nAPPENDIX A\nFig. 3.| Distribution of primary BH spin ( a1) { top panel; secondary BH spin ( a2) { middle panel; e\u000bective spin parameter ( \u001fe\u000b) {\nbottom panel; of BH-BH mergers at redshifts z <1:0. The results are for two tested models: the non-CE model plotted with red line\nand the CE model plotted with blue line. The \fgure is a supplement to the statistical spin predictions shown in Table 2 and described in\nSection 3." }, { "title": "1705.05395v1.Spin_and_tunneling_dynamics_in_an_asymmetrical_double_quantum_dot_with_spin___orbit_coupling.pdf", "content": "arXiv:1705.05395v1 [cond-mat.mes-hall] 15 May 2017Spin and tunneling dynamics in an asymmetrical double quant um dot with\nspin - orbit coupling\nMadhav Singh1, Pradeep K Jha2and Aranya B Bhattacherjee3\n1Department of Physics and Astrophysics, University of Delh i, Delhi-110007, India\n2Department of Physics, DDU College, University of Delhi, Ne w Delhi\n3School of Physical Sciences, Jawaharlal Nehru University, New Delhi-110067, India\nIn this articlewestudy the spin and tunneling dynamics asafunction o fmagneticfield in a\none-dimensionalGaAs double quantum dot with both the Dresselhau sand Rashbaspin-orbit\ncoupling. In particular we consider different spatial widths for the s pin-up and spin-down\nelectronic states. We find that the spin dynamics is a superposition o f slow as well as fast\nRabi oscillations. It is found that the Rashba interaction strength as well as the external\nmagnetic field strongly modifies the slow Rabi oscillations which is partic ularly useful for\nsingle qubit manipulation for possible quantum computer applications.\nI. INTRODUCTION\nSemiconductor quantum dots (QDs), also called “artificial a toms” are analogous to real atoms\n[1] and have been one of the most intensively studied nanostr uctures due to the richness in their\nproperties [2]. Numerous interesting transport propertie s of QD based systems have been revealed\nsuch as Coulomb blockade, spin blockade and Kondo effect [3–7] . One of the challenges in spintron-\nics and quantum transport is fast and efficient manipulation o f electron spins in QDs. Infact by\ncontrolling the electric bias, tunable spin diodes have bee n realized experimentally [8, 9]. Manipu-\nlated electron spins in QDs are expected as a possible realiz ation of qubits [10]. In the presence of\nspin-orbit (SO) coupling, single electron QDs exhibit spin -flip dynamics which can be understood\nin terms of the electric dipole spin resonance (EDSR) [11, 12 ]. Using EDSR, coherent spin manip-\nulation in GaAs QDs have been demonstrated, where electric fi eld induced spin Rabi oscillations\nhave been observed [13]. Other various SO effects present in se miconductor nanostructures provide\nan efficient and reliable way to manipulate electron spins in t wo-dimensional (2D) electron gases\n[14–21]. Manipulating spin degrees of freedom has opened ne w possibilities to design fast and low\npower quantum devices for applications in quantum computin g and memory storage [22, 23].\nHowever for quantum information applications, single isol ated QDs are not suitable since in-\nterdot interaction is necessary to produce and manipulate m any-body states. Double QDs where\ntunneling plays a crucial role are more promising for quantu m information technologies [24]. The2\n/MinuΣ4 /MinuΣ2 0 2 402468\nx/Slash1dV/LParen1x/RParen1\nFigure 1: Schematic plot of the double well potential described by Eq uation (1).\ninterplay between tunneling and spin flip process is an impor tant factor in double QDs. In the\npresence of SO interaction, the electron states and the inte rdot tunneling becomes spin-dependent.\nSingleelectrondynamicsinaone-dimensionaldoubleQDwit hSOcouplingdrivenbyanexternal\nelectric and magnetic fields has been a subject of earlier num erous study [25–28]. In these works,\nthe mutual effect of coordinate and spin motion on the Rabi osci llations were carried out under\nthe assumption that the spatial widths of the wacefunction f or spin-up and spin-down states are\nsame along the direction of the confining potential. However since the magnetic field lifts the\nspin degeneracy, the wavefunctions in the two spin-split le vels are not the same [29]. Keeping this\nimportant point in view, in this article we study the spin and tunneling dynamics as a function of\nmagnetic field in a one-dimensional double QD with SO couplin g and different spatial widths for\nthe spin-up and spin-down states. Both the Dresselhaus and R ashba SO coupling are taken into\naccount.\nII. THE MODEL AND HAMILTONIAN\nWe consider an electron confined in a one dimensional double Q D described by the quadratic\npotential as shown in figure1.\nV(x) =U0(−2x2\nd2+x4\nd4), (1)\nwhere the two minima located at d and -d are separated by a barr ier of height U0.\nThe unperturbed system is described by the Hamiltonian:3\nH0=ˆp2\nx\n2m+V(ˆx), (2)\nwhere ˆpx=−i/planckover2pi1∂\n∂xis the momentum operator in the x−direction. Here m is the mass of the\nelectron. In the absence of any external fields and spin-orbi t coupling (SOC), the ground state is\nsplit into the doublet of even (symmetric) and odd (anti symm etric) states. The tunneling energy\n∆Eg<< V0determines the tunneling time Ttun=2π\n∆Egand ∆Egalso is the gap between the\nground and first excited state of H0. Under the influence of an external magnetic field BzalongZ\ndirection and electric field along the x direction the system is described by Hamiltonian\nH=H0−eEx+HSO−∆zσz\n2. (3)\nThe second term −eExdescribe the perturbation caused by external electric field . The Zeeman\ncoupling to the magnetic field is∆zσz\n2, where ∆ z=|g|µBBzis the Zeeman splitting. The Pauli\nmatrices are described by σi=x,y,z. Heregis the Land´ e factor of electron and µBis the Bohr\nmagneton. The spin orbit (SO) interaction is described by Ha miltonian HSOwhich is the sum of\nthe bulk originated Dresselhaus( β) and structure related Rashba ( α) terms,\nHSO= (β\n/planckover2pi1pxσx−α\n/planckover2pi1pxσy). (4)\nIn order to study the spin and tunnelling dynamics as a functi on of the applied magnetic field ,\nwe use a perturbation approach and diagonalize the Hamilton ianHin the truncated spinor basis\nΨm(x)| ↑>with corresponding eigenvalues Em,σ. Here Ψ m(x) are the eigenfunctions of H0in the\ndouble well potential with m=s(symmetric), m=a(antisymmetric) and σ= +(−) corresponds\nto the spin parallel (antiparallel) to the zaxis. The wavefunctionΨ s(x)(Ψa(x)) is even (odd)\nwith respect to the inversion of x. Assuming weak tunneling, these symmetric and antisymmetric\nfunctions can be written in the form :\nΨs,a(x) =ΨL(x)±ΨR(x)√\n2, (5)\nwhere Ψ L(x) and Ψ R(x) are localized wavefunction in the left and right dot repect ively.\nΨL(x) =1√8wxsin(π\n2+nxπ\nwx), (6)4\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.5/Minus0.4/Minus0.3/Minus0.2/Minus0.10.0\nTime/LParen1Secs/RParen1/LeΣΣΣRz/Greatera\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus110.00.20.40.60.8\nTime/LParen1Secs/RParen1PR/LParen1t/RParen1b\nFigure 2: Dynamics of σz\nR(t) (a) and PR(t) (b) for the identical width (ID) case for B= 0.88T. The various\nparameters used are: α= 1×10−9eV−cm,β= 0.3×eV−cm, g=-0.45, wx= 25√\n2nm,wx=w′\nx\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.6/Minus0.4/Minus0.20.00.2\nTime/LParen1Secs/RParen1/LeΣΣΣRz/Greatera\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus110.400.450.500.550.600.65\nTime/LParen1Secs/RParen1PR/LParen1t/RParen1b\nFigure 3: Dynamics of σz\nR(t) (a) and PR(t) (b) for the identical width (ID) case for B= 6.4T. The various\nparameters used are: α= 1×10−9eV−cm,β= 0.3×eV−cm, g=-0.45, wx= 25√\n2nm,wx=w′\nx.\nΨR(x) =1√8wxsin(π\n2−nxπ\nwx). (7)\nHerewxis the spatial width of the wavefunction. We will be restrict ing our analysis to\ntwo lowest states n= 1,2. The first four lowest energy states are |Ψ1>= Ψs(x)| ↑>,|Ψ2>=\nΨs(x)| ↓>,|Ψ3>= Ψa(x)| ↑>and|Ψ4>= Ψa(x)| ↓>. The spatial parts of the wavefunctions\nΨs(a)(x) are different for spin-up and spin-down states. The spin dege neracy is lifted due to the\nmagnetic field and as a result of which the wavefunctions in th e two spin split levels are not the\nsame. Hence the spatial spread of the wavefunctions are differ ent . The wavefunction correspond-\ning to the state with higher energy spreads out more outside t he potential well (spatial width w′\nx\n) compared to the wavefunction corresponding to the state wi dth lower energy level (spatial width\nwx) andw′\nx> wx. The full dynamics of the system can then be studied with the f unction5\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.4/Minus0.20.00.20.4\nTime/LParen1Secs/RParen1/LeΣΣΣRz/Greatera\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus110.40.50.60.7\nTime/LParen1Secs/RParen1PR/LParen1t/RParen1b\nFigure 4: Dynamics of σz\nR(t) (a) and PR(t) (b) for the non-identical width (NID) case for B= 0.88T.\nThe various parameters used are: α= 1×10−9eV−cm,β= 0.3×eV−cm, g=-0.45, wx= 25√\n2nm,\nw′\nx= 1.0375wx.\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.4/Minus0.20.00.20.4\nTime/LParen1Secs/RParen1/LeΣΣΣRz/Greatera\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus110.350.400.450.500.550.600.65\nTime/LParen1Secs/RParen1PR/LParen1t/RParen1b\nFigure 5: Dynamics of σz\nR(t) (a) and PR(t) (b) for the non-identical width (NID) case for B= 6.4T.\nThe various parameters used are: α= 1×10−9eV−cm,β= 0.3×eV−cm, g=-0.45, wx= 25√\n2nm,\nw′\nx= 1.0375wx.\n|Ψ>=/summationdisplay\nnξn(t)e−iEnt\n/planckover2pi1|Ψn>, (8)\nwhereξn(t) are the expansion coefficients . Once the wavefunctions are k nown, we can calculate\nthe probability PR(t) to find the electron in the right dot and expectation value of theithspin\ncomponent σi\nRin the right dot since we are interested in the quantum transp ort from the left to\nthe right quantum dot.\nPR(t) =/integraldisplay∞\n0Ψ†Ψdx, (9)6\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.4/Minus0.20.00.20.40.6\nTime/LParen1Secs/RParen1/LeΣΣΣRx/Greatera\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.6/Minus0.4/Minus0.20.00.20.40.6\nTime/LParen1Secs/RParen1/LeΣΣΣRy/Greaterb\nFigure 6: Dynamics of σx\nR(t) (a) and σy\nR(t) (b) for the non-identical width (NID) case for B= 0.88T.\nThe various parameters used are: α= 1×10−9eV−cm,β= 0.3×eV−cm, g=-0.45, wx= 25√\n2nm,\nw′\nx= 1.0375wx.\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.6/Minus0.4/Minus0.20.00.20.40.6\nTime/LParen1Secs/RParen1/LeΣΣΣRx/Greatera\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.6/Minus0.4/Minus0.20.00.20.40.6\nTime/LParen1Secs/RParen1/LeΣΣΣRy/Greaterb\nFigure 7: Dynamics of σx\nR(t) (a) and σy\nR(t) (b) for the non-identical width (NID) case for B= 6.4T.\nThe various parameters used are: α= 1×10−9eV−cm,β= 0.3×eV−cm, g=-0.45, wx= 25√\n2nm,\nw′\nx= 1.0375wx.\nσi\nR(t) =/integraldisplay∞\n0Ψ†σiΨdx. (10)\nThe explicit expressions for PR(t) andσi\nR(t) are given in the appendix A.\nIII. RESULT AND DISCUSSION\nWe are interested in the inter dot transition. To demonstrat e the nontrivial dynamics of the\ntunnelling and spin flip process, we plot the dynamics of PR(t) andσi\nR(t)(i=x,y,z) for different\nmagnetic fields.\nFor the ID case as shown in Fig.2 and Fig.3, the spin component σz\nR(t) displays an irregular7\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.3/Minus0.2/Minus0.10.00.10.20.3\nTime/LParen1Secs/RParen1/LeΣΣΣRx/Greatera\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.2/Minus0.10.00.10.2\nTime/LParen1Secs/RParen1/LeΣΣΣRy/Greaterb\nFigure 8: Dynamics of σx\nR(t) (a) and σy\nR(t) (b) for the non-identical width (NID) case for B= 11T.\nThe various parameters used are: α= 1×10−9eV−cm,β= 0.3×eV−cm, g=-0.45, wx= 25√\n2nm,\nw′\nx= 1.0375wx.\n0 2 4 6 8 102.662.682.702.72\nB/LParen1Tesla/RParen1/VertBar1/CΑpDeltΑ1/Plus/CΑpDeltΑ2/VertBar1/LParen1meV/RParen1a\n0 2 4 6 8 100.080.100.120.140.16\nB/LParen1Tesla/RParen1/VertBar1/CΑpDeltΑ1/Minus/CΑpDeltΑ2/VertBar1/LParen1meV/RParen1b\nFigure 9: Plot of ∆(1)+∆(2)(a) and ∆(1)−∆(2)(b) as a function of the magnetic field .\npattern and the mean value of σz\nR(t) (denoted by ( σz\nR(t))M) increases from −0.10 atBz= 11Tto\n−0.25 atBz= 0.88T. Similarly the dynamics of PR(t) is also irregular and the mean value of\nPR(t)(denoted by ( PR(t))M) increases from 0 .35 atBz= 11Tto 0.52 at Bz= 0.88T. The strong\noscillations of PR(t) around a mean value of 0 .52 for small and moderate fields Bz= 0.88T(Fig.\n2b),Bz= 2.82T(not shown) and Bz= 6.4T(Fig. 3b), indicates that the external field almost\nequalizes the probabilities to find the electron in the right and left dot . For large magnetic field of\n11T, the mean value around which PR(t) oscillates is 0 .35 (not shown) which shows that a strong\nmagnetic field of 11 Tsuppresses the tunneling from the left to the right well. Thi s indicates that\nfor small magnetic fields, a small spin polarization can be ac hieved in the tunneling under ID case.\nThe dynamics for the NID case is more regular and rich as depic ted in figure 4 and figure 5.\nThe dynamics comprises of low frequency motion due to the ter m ∆(1)−∆(2)superimposed by\nhigh frequency oscillations due to the involvement of term ∆(1)+∆(2)in the expression for PR(t)\nandσi\nR(t) (see appendix). For the NID case, ( σz\nR(t))Mis zero for all magnetic fields while ( PR)M8\n01./Multiply10/Minus112./Multiply10/Minus113./Multiply10/Minus114./Multiply10/Minus115./Multiply10/Minus116./Multiply10/Minus11/Minus0.4/Minus0.20.00.20.4\nTime/LParen1Secs/RParen1/LeΣΣΣRz/Greatera\n0 1 2 3 4 50246810\nΑ/LParen1/Multiply10/Minus9eV/Minuscm/RParen1/VertBar1/CΑpDeltΑ1/Minus/CΑpDeltΑ2/VertBar1/LParen1meV/RParen1,BZ/EquΑl0.88 Tb\nFigure10: Plot(a): Dynamicsof σx\nR(t)(a)forthenon-identicalwidth(NID) casefor B= 0.88T. Thevarious\nparameters used are: α= 2×10−9eV−cm,β= 0.3×eV−cm, g=-0.45, wx= 25√\n2nm,w′\nx= 1.0375wx.\nPlot (b): ∆(1)−∆(2)versus Rashba interaction strength ( α) for the NID case.\nincreases from 0 .32 atBz= 11Tto 0.54 atBz= 0.88T, which indicates that no spin- polarization\ncan be achieved in the tunneling.\nThe SO interaction is the sum of bulk originated Dresselhaus (β) and structure related Rashba\n(α) terms. Inthetunneling, spinprecession aroundboththe xas well as yaxis isobserved as shown\nin figure 6 and figure 7 for the NID case. The spin precession aro und the y axis occurs due to the\nRashbaSOterm whilethespinprecession aroundthe xaxis occursduetotheDresselhausSOterm.\nThe magnetic field produces the zeeman spin splitting of the l evel ∆ z=En↓−En↑=|g|µBBzfor\nthe ID case. On the other hand for the NID case, the spin splitt ing energy is given by [29],\n∆z= 2/radicalBigg/bracketleftbigggµBB\n2/bracketrightbigg2\n+α264\nwxw′\nxf(wx,w′\nx), (11)\nwhere\nf(wx,w′\nx) = cos2(πwx\n2w′\nx)cos2(πw′\nx\n2wx)/bracketleftbigg1\n((wx/w′\nx)2−1)((w′\nx/wx)2−1)/bracketrightbigg2\n. (12)\nFor the NID case, we note that the spin split energy is larger a s compared to that for ID case.\nWith the increase in magnetic field , the effect of SOC decreases , leading to smaller amplitudes\nof spin precession as can be seen by comparing figures 6 and 7 wi th figure 8. Comparing the\ndynamics of < σz\nR(t)>from Figs. 4 and Fig. 5, there is a minor increase in the Rabi fr equency\n|∆(1)+∆(2)|with increasing magnetic fieldwhich is consistent with prev ious theoretical [11, 12] and\nexperimental result [13]. On the other hand the slow oscilla tions|∆(1)−∆(2)|becomes slower with9\nincreasingmagneticfield. Thesetwoobservationsareclear ly depictedinFig.9where∆(1)+∆(2)and\n∆(1)−∆(2)areplotted as a function of the magnetic field in Fig.9a and Fi g.9b respectively. Thelow\nfrequency dynamics corresponds to the spectrum of the low en ergy states while the high frequency\noscillations involve the higher-energy states. As evident from Figures 6,7 and 8, increasing the\nmagnetic field, the slow oscillations of < σx\nR(t)>and< σy\nR(t)>gradually disappear at Bz= 11T.\nThus at high magnetic fields, only the higher energy states ar e involved in the dynamics. The\nsame conclusion cannot be reached for the ID case due to its ir regular dynamics. In addition, we\nobserve that for the ID case, both the initial state and spin p recession axis change with magnetic\nfield leading to different phase shifts between the spin compon ents. On the other hand, for the\nNID case, the initial state and the spin precession axis rema ins intact for different magnetic fields.\nThe incomplete Rabi spin flips results from the fact that the t he electron tunneling between the\npotential minimas establishes a competing spin dynamics wh ich prevents the electric field to flip\nthe spin efficiently ( σz\nR(t) to reach ±1). The slow tunneling dynamics of the NID case (Fig.4b and\n5b) allows the electron to stay in the right well for a longer d uration. The spin precession in this\nslow interminima motion induces a corresponding spin dynam ics which does not allow the electric\nfield to flip the spin efficiently and thus spin polarization is a bsent for the NID case. This is in\ncontrast to the ID case where the tunneling dynamics is extre mely fast and the electron oscillates\nrapidly between the left and the right well. The fast inter-m inima motion washes out the induced\nspin dynamics and thus allows the electric field to flip the spi n.\nSpintronicsbasedimplementation ofscalablequantumcomp utershasgeneratedmuchinterest in\nexploringcoherentcontrolofsinlequbitrotationusingRa bioscillations[29]. Inordertosuccessfully\nimplement single qubit rotation selectively in a sample com prising of multiple quantum dots using\nRabi oscillations, thespin splitting energy in the target q uantum dot has to bechanged appreciably\nusing Rashba interaction [29]. It was shown in ref. [29] that the energies of the spin-split levels\nof the lowest subband in a QD decreases with increasing Rashb a interaction strength. Since Rabi\noscillations play an important role in the control of single qubit rotation, we have studied here\nthe dynamics of < σz\nR(t)>as a function of Rashba interaction strength α. Fig.10a shows the\nplot of< σz\nR(t)>as a function of time for α= 2×10−9eV−cmandB= 0.88T. Clearly the\nfrequency of the slow oscillations |∆(1)−∆(2)|increases as compared to that in Fig. 4a where\nα= 1×10−9eV−cm. This is also depicted in Fig.10b where |∆(1)−∆(2)|is seen to increase\nwithα. The influence of αon|∆(1)+∆(2)|is not appreciable. These observations indicates that\nincreasing the Rashba interaction strength induces a rapid energy transfer between the low energy\nstates while energy transfer between the higher energy stat es essentially remains uneffected.10\nIV. CONCLUSION\nIn conclusion, we have studied the quantum spin and charge dy namics of a single electron\nconfined in one-dimensional GaAs double quantum dot with ras hba and Dresselhaus spin-orbit\ninteraction. We have made a comparitive study of two specific cases: In first case, the spatial\nparts of the wavefunctions are same for spin-up and spin-dow n states (ID case) while in the second\ncase we consider the spatial parts of the wavefunctions to be different for spin-up and spin-down\nstates (NID case). We demonstrate that the dynamics of the NI D case is more regular. The\ndynamics comprises of a slow as well as fast Rabi oscillation s. The slow oscillations corresponds to\ndynamics involving low energy states while fast oscillatio ns involve the higher energy states. For\nthe ID case, we found a small amount of spin polarization in th e tunneling but for the NID case, a\ncomplete absence of spin polarization in the tunneling is no ticed. We found that a coherent control\nof the slow Rabi oscillations is possible using the Rashba in teraction strength and the external\nmagnetic field which makes this scheme useful for single qubi t manipuation for quantum computer\napplications.\nAcknowledgments\nA. B acknowledges financial support from the University Gran ts Commission, New Delhi under\nthe UGC-Faculty Recharge Programme. M. S acknowledges finan cial support from the University\nGrants Commission for the junior research fellowship.11\nV. APPENDIX A\nIdentical Width Case\nσy\nR(t) =N2(0.84(a1b2−a2b1)sin(2∆(1)t)+0.998(d2a1−a2d1+b2c1−b1c2)sin([∆(1)+∆(2)]t)\n+ 0.998(c2a1−a2c1+b2d1−b1d2)sin([∆(1)−∆(2)]t)+0.84(c1d2−c2d1)sin(2∆(2)t);\nσz\nR(t) =N2(1\n2(a2\n2−a2\n1+b2\n2−b2\n1+c2\n2−c2\n1+d2\n2−d2\n1)+(a2b2−a1b1)cos(2∆(1)t)\n+ (c2d2−c1d1)cos(2∆(2)t)+8\n3π(a2c2−a1c1)cos([∆(1)−∆(2)]t)\n+8\n3π(a2d2−a1d1+b2c2−b1c1)cos([∆(1)+∆(2)]t));\nσx\nR(t) =N2(0.858(a1a2+b1b2+c1c2+d1d2)+0.858(a2b1+a1b2)cos(2∆(1)t)\n+ 0.858(d1c2+c1d2)cos(2∆(2)t)+0.998(a2d1+a1d2+b2c1+b1c2)cos([∆(1)+∆(2))]t)\n+ 0.998(a2c1+a1c2+b2d1+b1d2)cos([∆(1)−∆(2)]t);\nPR(t) =N2(1\n2(a2\n2+a2\n1+b2\n2+b2\n1+c2\n2+c2\n1+d2\n2+d2\n1)+8\n3π(a2d2+a1d1+b2c2+b1c1)cos([∆(1)+∆(2)]t)\n+8\n3π(b2d2+b1d1+a1c1+a2c2)cos([∆(1)−∆(2)]t)+(a1b1+a2b2)cos(2∆(1)t)\n+ (c1d1+c1d2)cos(2∆(2)t), (13)12\nNon-identical width case\nσx\nR(t) =N2(8\n3π(a1a2+b1b2+c1c2+d1d2)+8\n3π(a2b1+a1b2)cos(2∆(1)t)\n+8\n3π(c2d1+c1d2)cos(2∆(2)t)+(a2c1+a1c2+b2d1+b1d2)cos(∆(1)t);\nσy\nR(t) =N2(8\n3π(a1b2−a2b1sin(2∆(1)t)+8\n3π(c1d2−c2d1sin(2∆(2)t)\n+ (a1c2−a2c1+b2d1−b1d2)sin([∆(1)−∆(2)]t)+(a1d2−a2d1+b2c1−b1c2)sin([∆(1)+∆(2)]t);\nσz\nR(t) =N2(1\n2(a2\n2−a2\n1+b2\n2−b2\n1+c2\n2−c2\n1+d2\n2−d2\n1)+(a2b2−a1b1)cos(2∆(1)t)\n+ (c2d2−c1d1)cos(2∆(2)t)+8\n3π(a2d2−a1d1+b2c2−b1c1)cos([∆(1)+∆(2)]t)\n+8\n3π(a2c2−a1c1)cos([∆(1)−∆(2)]t);\nPR(t) =N2(1\n2(a2\n1+a2\n2+b2\n1+b2\n2+c2\n1+c2\n2+d2\n1+d2\n2)+8\n3π(a2d2+a1d1+b2c2+b1c1)cos([∆(1)+∆(2)t)\n+8\n3π(b2d2+b1d1+a1c1+a2c2)cos([∆(1)−∆(2)]t)+(a1b1+a2b2)cos([2∆(1)]t)\n+ (c1d1+c2d2)cos(∆(2))t). (14)13\n|a1|2=(−eEη+(/radicalbig\nα2+β2)sin(πw′\nx\n2wx)2√\nwxw′\nx\n4w2x−(w′\nx)2)2\n(∆E1\n2+∆z\n2+∆(1))2+(−eEη+(/radicalbig\nα2+β2)sin(πw′\nx\n2wx)2√\nwxw′\nx\n4w2x−(w′\nx)2)2; (15)\n|a2|2= 1− |a1|2; (16)\n|b1|2=(−eEη+(/radicalbig\nα2+β2)sin(πw′\nx\n2wx)2√\nwxw′\nx\n4w2x−(w′\nx)2)2\n(−∆E1\n2−∆z\n2+∆(1))2+(−eEη+(/radicalbig\nα2+β2)sin(πw′\nx\n2wx)2√\nwxw′\nx\n4w2x−(w′\nx)2)2; (17)\n|b2|2= 1− |b1|2; (18)\n|c1|2=(−eEη1−(/radicalbig\nα2+β2)cos(πw′\nx\nwx)2√\nwxw′\nx\nw2x−4(w′\nx)2)2\n(∆E2\n2−∆z\n2−∆(2))2+(−eEη1−(/radicalbig\nα2+β2)cos(πw′\nx\nwx)2√\nwxw′\nx\nw2x−4(w′\nx)2)2; (19)\n|c2|2= 1− |c1|2; (20)\n|d1|2=(−eEη1−(/radicalbig\nα2+β2)cos(πw′\nx\nwx)2√\nwxw′\nx\nb2−4(w′\nx)2)2\n(∆E2\n2−∆z\n2+∆(2))2+(−eEη1−(/radicalbig\nα2+β2)cos(πw′\nx\nwx)2√\nwxw′\nx\nw2x−4(w′\nx)2)2; (21)\n|d2|2= 1− |d1|2; (22)\nη=−64wxw′\nxcos(πw′\nx\nwx)+8πsin(πw′\nx\n2wx)(4w2\nx−(w′\nx)2\nπ2((w′\nx)2−4w2x)21\n(wxw′\nx)3\n2; (23)\nη1=64wxw′\nxsin(πw′\nx\n2wx)+16πcos(πw′\nx\nwx)(w2\nx−4(w′\nx)2\nπ2(−4(w′\nx)2+w2x)21\n(wxw′\nx)3\n2; (24)\nN=/radicalBigg\n1\n4+2(a1b1+a2b2+c2d1+c2d2). (25)\n∆(1)=1\n2/radicalbig\n(∆E1+∆z)2+4c2, (26)\n∆(2)=1\n2/radicalBig\n(∆E2−∆z)2+4(c′)2, (27)\nc=−eEη+/radicalbig\nα2+β22√\nbb′\n4b2−b′2sinπb′\n2b, (28)\nc′=−eEη1−/radicalbig\nα2+β22√\nbb′\n−4b2+b′2cosπb′\nb, (29)14\nHere ∆E1=E′\n2−E1and ∆E2=E2−E′\n1, where E1andE2are the lowest energies of ↑\nstate while E′\n1andE′\n2are the lowest energies of ↓state. Here a1,a2,b1,b2,c1,c2,d1andd2are the\neigenvalues.\n[1] J. P. Bird, Electron transport in Quantum Dots, (Spinger, Berlin , 2003).\n[2] S. Kohler, J. Lehmann and P. Hanggi, Phys. Rep. 405, 379 (2005).\n[3] P. Recher, E. V. Sukhorukov and D. Loss, Phys. Rev. Lett. 85, 1962 (2000).\n[4] D. Weinmann, W. Hausler and B. Kramer, Phys. Rev. Lett., 74, 984 (1995).\n[5] S. Takahashi and S. Maekawa, Phys. Rev. Lett., 80, 1758 (1998).\n[6] W. Liang et al., Nature (London), 417, 725 (2002).\n[7] J. Park et al., Nature (London), 417, 722 (2002).\n[8] A. Merchant and N. Markovi ´ c, Phys. Rev. Lett., 100, 156601 (2008).\n[9] K. Hamaya et al., Phys. Rev. Lett., 102, 236806 (2009).\n[10] G. Burkard, D. Loss and D. P. DiVincenzo, Phys. Rev. B., 59, 2070 (1999).\n[11] E. L. Rashba and AI. L. Efros, Phys. Rev. Lett. 91, 126405 (2003).\n[12] E. L. Rashba and AI. L. Efros, Appl. Phys. Lett. 83, 5295 (2003).\n[13] K. C. Nowack, F. H. L. Koppens, Yu. V. Nazarovand L. M. K. Va ndersypen, Science 318, 1430(2007).\n[14] R. Winkler, Spin-Orbit coupling effects in two-dimensional electro n and hole systems, Springer tracts\nin modern physics, 191, (Springer, Berlin, 2003).\n[15] Y.V.Pershinnd C. Piermarocchi, Phys. Rev. B 72, 245331 (2005).\n[16] J. S. Sheng and K. Chang, Phys. Rev. B 74, 235315 (2006).\n[17] E. Rasanen, A. Castro, J. Werschnik, A. Rubio, and E. K. U. Ph ys.Rev. Lett. 98, 157404 (2007).\n[18] W.-H. Kuan, C.-S. Tang, and C.-H. Chang, Phys. Rev. B 75, 155326 (2007).\n[19] D.-Y. Liu and J.-B. Xia, J. Appl. Phys. 115, 044313 (2014).\n[20] F. K. Joibari, Ya. M. Blanter, and G. E. W. Bauer, Phys. Rev. B 90, 155301 (2014).\n[21] S. Gumber,M. Gambhir, P. K. Jha, and Man Mohan J. Appl. Phys. 119, 073101 (2016).\n[22] S. A. Wolf et al., Science, 294, 1488 (2001).\n[23] Semiconductor spintronics and quantum computation, editied b y D. D. Awschalom, N. Samarth and\nD. Loss (Springer, Berlin, 2002).\n[24] J. R. Petta et al., Science, 309, 2180 (2005).\n[25] D. V. Khomitsky, L. V. Gulyaev and E. Ya. Sherman Phys. Rev. B 85, 125312 (2012).\n[26] D. V. Khomitsky and E. Ya. Sherman, Phys. Rev. B 79, 245321 (2009).\n[27] D. Khomitsky and E. Sherman, Nanoscale Research Letters 6, 212,(2011).\n[28] D. V. Khomitsky and E. Ya. Sherman, EPL, 90, (2010) 27010.\n[29] D. Bhowmik and S. Bandyopadhyay, Physica E, 41, 587 (2009)." }, { "title": "2005.01391v1.Spin_orbit_torque_magnonics.pdf", "content": "1\nSpin-orbit-torque magnonics \nV. E. Demidov,1* S. Urazhdin,2 A. Anane,3 V. Cros,3 S. O. Demokritov1 \n1Institute for Applied Physic s and Center for Nanotechnology , University of Muenster, \nCorrensstrasse 2-4, 48149 Muenster, Germany \n2Department of Physics, Emory University, Atlanta, GA 30322, USA \n3Unité Mixte de Physique, CNRS, Thales, Uni versité Paris-Saclay, 91767, Palaiseau, France \nThe field of magnonics, which ut ilizes propagating spin waves f or nano-scale transmission \nand processing of information, has been significantly advanced by the advent of the spin-orbit \ntorque. The latter phenomenon ca n allow one to overcome two mai n drawbacks of magnonic \ndevices – low energy efficienc y of conversion of electrical sig nals into spin wave signals, and \nfast spatial decay of spin waves in thin-film waveguiding struc tures. At first glance, the \nexcitation and amplification of s pin waves by spin-orbit torque s can seem to be \nstraightforward. Recent researc h indicates, however, that the lack of the mode-selectivity in \nthe interaction of spin current s with dynamic magnetic modes an d the onset of dynamic \nnonlinear phenomena represent si gnificant obstacles. Here, we d iscuss the possible route to \novercoming these limitations, based on the suppression of nonli near spin-wave interactions in \nmagnetic systems with perpendicular magnetic anisotropy. We sho w that this approach \nenables efficient excitation of coherent magnetization dynamics and propagating spin waves \nin extended spatial regions, a nd is expected to enable practica l implementation of complete \ncompensation of spin-wave propagation losses. \nKeywords: spin current, ma gnonics, magnetic nanostructures \n* Corresponding author, e-mail: demidov@uni-muenster.de \n 2\nI. Introduction \nHigh-frequency spin waves propa gating in thin ferromagnetic fil ms have been utilized for \nthe implementation of advanced microwave devices for many decad es.1-5 Unique features of \nthese waves include electronic tunability by the magnetic field , very short (millimeter to \nsubmicrometer) wavelengths in the microwave frequency range, as well as controllability of \npropagation characteristics by the direction of the static magn etic field relative to the direction \nof wave propagation.3-5 These features made spin waves very attractive for the impleme ntation \nof a variety of devices for comm unication technologies, such as microwave filters, phase \nshifters, delay lines, multiplexers, etc. With the rapid increa se in recent years of the operational \nfrequencies of conventional elec tronic systems, in particular t hose for computation and \ninformation processing, high-fre quency spin waves have become i ncreasingly attractive for the \nimplementation of integrated logic circuits and nanoscale wave- based computing systems.6-8 \nThis has motivated the emergence of a new research field - magn onics,9-12 - which explores the \npossibility to utilize spin wav es for transmission and processi ng of information at nanoscale as \nan alternative to conventional CM OS-based electronic circuits. \nDespite numerous advantages provided by spin waves, magnonic de vices currently suffer \nfrom two drawbacks. First, the inductive mechanism traditionall y utilized to convert electrical \nsignals into spin waves and back, is characterized by relativel y low conversion efficiency.13-18 \nEven if low conversion losses can be achieved in traditional ma croscopic spin-wave devices \nwith millimeter-scale dimensions, miniaturization of magnonic d evices down to the sub-\nmicrometer scale inevitably results in large conversion losses unacceptable for real-world \napplications. The second drawback of microscopic magnonic devic es is associated with large \npropagation losses of spin waves in nanometer-thick magnetic fi lms. In macroscopic spin-wave \ndevices based on high-quality micro meter-thick films of Yttrium Iron Garnet (YIG), the decay \nlength of spin waves reaches several millimeters. However, in m icroscopic waveguides 3\npatterned from ultrathin YIG f ilms, the decay length typically does not exceed a few tens on \nmicrometers19-21. \nTo make the microscopic magnonic devices technologically compet itive, it is necessary to \novercome their two main drawback s described above. This may bec ome possible th anks to the \nadvent of spin-torque phenomena22,23, which allow reduction of magnetic damping, and even \nits complete compensation, enabling excitation of high-frequenc y magnetization oscillations by \ndc electrical currents24,25. In particular, it has been show n that spin torques created by spin-\npolarized electric currents or by non-local spin injection can excite propagating spin waves26-\n32, and that this excit ation mechanism is intrinsically fast31 and highly power-efficient.30 These \nachievements clearly demonstrated a large potential of spin tor ques for magnonics. However, \nthe devices proposed in Refs. 26-32 utilized current injection through a magnetic nanocontact , \nresulting only in local spin torque. It could not provide compe nsation of damping over spatially \nextended regions, which is of vital importance for magnonic app lications that require enhanced \nspin wave propagation. \nIn contrast to the conventional spin torques produced by the lo cal current injection through \nconducting magnetic materials, the spin-orbit torque (SOT) asso ciated with charge to spin \nconversion in the bulk or at interfaces in material systems wit h strong spin-orbit interaction,33-\n36 provides the ability to control magnetic damping in spatially extended regions of both \nconducting and insulating magnetic materials37-39. This may enable decay-free propagation of \nspin waves, or even their true amplification in magnetic nanost ructures. In recent years, this \npossibility was intensively explored for macroscopic40-43 and microscopic44-48 magnonic \ndevices. The largest effect reported so far was achieved in mic roscopic magnonic waveguides \nbased on ultrathin YIG films,46 where a nearly tenfold increase in the propagation length of \ncoherent spin waves was demonstr ated. The high efficiency of th e system studied in Ref. 46 \npermitted experimental access to the operational regime close a nd even above the point at which 4\ndamping is expected to be completely compensated by SOT. Contr ary to naive expectations, it \nwas found that the onset of nonlinear damping in the overdriven magnon system does not allow \none to achieve complete compensation of spin-wave propagation l osses, required for true \namplification of cohere nt spin waves by SOT. \nNote that the same fundamental mechanisms also prevent SOT-indu ced excitation of \ncoherent magnetization oscillati ons in spatially extended syste ms.39,49 In Ref. 50, it was shown \nthat the adverse effects of non linear damping can be suppressed by localizing the injection of \nspin current to a nanoscale regi on of the magnetic system. This a p p r o a c h e n a b l e d t h e \nimplementation of SOT-driven ma gnetic nano-oscillators that exh ibit coherent high-frequency \nmagnetization oscillations50-61. However, it also imposed strict limitations on the layout of these \ndevices, hindering their utiliz ation as sources of propagating spin waves. As a result, it took a \nsignificant time to understand how confined-area SOT-driven osc illators can be integrated with \na medium supporting spin-wave propagation.62,63 We emphasize that, while spin-wave \nexcitation by SOT was demonstrat ed in Refs. 62,63, these works have not resolved the general \nproblem of the adverse eff ects of nonlinear damping. \nIn this brief perspective article, we present our vision of the main outstanding issues \nhindering further developments in the field of SOT-driven magno nics, and outline an approach \nthat can help to overcome them. We first discuss the key recent experimental results revealing \nthe physical nature of these issues. In particular, we show tha t the functionality of SOT \nmagnonic devices is strongly limite d by the lack of spin-wave m ode selectivity of the \ninteraction of spin currents with the magnetization, which resu lts in the simultaneous \nenhancement by SOT of many incoherent spin-wave modes. This enh ancement causes an onset \nof the nonlinear scattering of coherent spin waves, which count eracts their enhancement by the \nanti-damping effect of SOT. We discuss a recently proposed appr oach that allows efficient \ncontrol of nonlinear damping, by utilizing reduction of ellipti city of magnetization precession 5\nin magnetic films where the dipolar anisotropy is compensated b y the interfacial perpendicular \nmagnetic anisotropy (PMA). This approach has been recently show n to enable SOT-driven \nexcitation of coherent magnetization auto-oscillations in spati ally extended systems based on \nconductive64 and insulating65 magnetic materials. It was also demonstrated to enable efficie nt \nexcitation of coherent propagating spin waves in magnetic insul ators.65 We project that this \napproach can also enable the implementation of decay-free propa gation of spin waves, \nresolving, in this way, the main issues limiting the progress i n nano-magnonics. \nII. Amplification of spin waves by SOT \nFigure 1 presents the results of the experiment that demonstrat ed highly efficient \ncompensation of propagation losses of spin waves by SOT46. The test devices used in this study \n(Fig. 1(a)) are 1 m-wide spin-wave waveguides patterned from a 20 nm-thick YIG fi lm \ncovered by an 8 nm-thick layer of Pt. Propagating spin waves ar e excited in the waveguide by \nmicrowave current flowing through a 3 m wide and 250 nm thick Au inductive antenna. \nAdditionally, a dc electrical current I is applied through the Pt layer of the waveguide. The \nelectrical current is converted into an out-of-plane pure spin current IS (see the inset in Fig. \n1(a)), due to the spin-Hall effect (SHE)33-36 in Pt. In turn, the spin current exerts an anti-damping \nspin-transfer torque37 on the magnetization M in YIG, which is expected to compensate the \npropagation losses of the coherent spin wave emitted by the ant enna. \nThe effects of SOT on the spin-w ave propagation were studied by using micro-focus \nBrillouin light scattering (BLS) spectroscopy.66 The probing laser light was focused through \nthe sample substrate into a diffraction-limited spot on the YIG /Pt film (Fig. 1(a)), and the \nmodulation of light due to its interaction with magnetic excita tions in YIG was analyzed. The \nresulting signal – the BLS inten sity – is proportional to the i ntensity of magnetic oscillations at \nthe position of the probing spot, which enables mapping of the intensity of propagating spin \nwaves with high spatial r esolution (Fig. 1(b)). 6\nFigure 1(c) shows several represen tative dependences of spin-wa ve intensity on the \npropagation coordinate x, obtained at different dc currents in the Pt layer. As seen fr om these \ndata, spin waves in the waveguide experience well-defined expon ential decay ~ exp(-2 x/), \nwhere is the decay length, defined as the distance over which the wa ve amplitude decreases \nby a factor of e. The decay length i ncreases with the increase of the dc current, as expected for \nthe effects of SOT on the effective magnetic damping. Figure 1( d) shows the current \ndependences of the decay length and of its inverse value – the decay constant, which is \nproportional to the effective Gilbert damping constant. The dec ay length monotonically \nincreases at small currents, reaches a maximum at a certain cur rent I = IC, and then abruptly \ndecreases at larger currents. Independent measurements were use d to identify IC as the critical \ncurrent, at which SOT is expected to completely compensate the natural damping in YIG. Thus, \none could expect that diverges at I = IC (the decay constant vanishes), and that the propagating \nwave becomes spatially amplified at I > IC. These naive expectations are clearly inconsistent \nwith the experimental findings. \nThis experiment demonstrated tha t SOT is capable to increasing the propagation length of \nspin waves by nearly an order of magnitude. For the studied sys tem, this corresponded to the \nincrease of the spin-wave intensity at the output of a 10 m-long transmission line by three \norders of magnitude. Simultaneously, this experiment showed tha t the simple picture of SOT \neffects on the damping of coherent propagating spin waves becom es invalid in the vicinity of \nthe point of the complete damping compensation. As will be disc ussed below, in the regime of \nlarge currents, it is necessary t o take into account not only t hat SOT reduces the effective \ndamping, but also that it strongly enhances incoherent magnetic fluctuations. The nonlinear \nscattering of coherent waves from these fluctuations represents an additional damping channel \ncounteracting the anti- damping effect of SOT. \n 7\nIII. Excitation of spin waves by SOT \nWhile SOT-induced coherent magne tic auto-oscillations have been achieved in several \nnanomagnetic device geometries50-53, and many other novel SOT oscillators have been \nproposed in recent years54-61, none of them provided the possibility to generate coherent \npropagating spin waves. This is mainly associated with the limi tations imposed by the device \nlayout necessary to achieve coherent auto-oscillations. Indeed, as was shown in Refs. 39,49, \ncoherent oscillations cannot be excited in spatially extended s ystems with arbitrary shape, due \nto the nonlinear interactions with incoherent spin-wave modes s trongly enhanced by SOT. To \nsuppress these nonlinear effects, spin current must be injected into a nanoscale region of the \nmagnetic system, which imposes strict limitations on the layout of SOT-driven auto-oscillators. \nAfter an intensive search for a suitable geometry, in Ref. 62, we proposed a new concept \nof nano-notch SOT auto-oscillato rs directly incorporated into a magnonic nano-waveguide. \nThese devices (Fig. 2(a)) were based on 180 nm-wide nano-wavegu ides patterned from a \nPermalloy(Py)(15 nm)/Pt(4 nm) bi layer. Ion milling was used to pattern a rectangular 200 nm-\nwide and 10 nm-deep notch in the top Py layer of the waveguide, forming a nano-oscillator that \nserves as the spin-wave source. When electric current I flows through the device, SHE in Pt \ninjects pure spin current IS into the Py layer, producing anti-damping SOT acting on its \nmagnetization M. The thickness-averaged magnitude of the anti-damping torque i s inversely \nproportional to the thickness of the magnetic layer. Thus, the SOT effects on the 5 nm-thick Py \nlayer in the nano-notch area ar e significantly larger than on t he 15 nm-thick Py waveguide. As \nthe current I is increased, damping becomes completely compensated in the na no-notch region, \nresulting in the local excitat ion of magnetizati on auto-oscilla tions. \nThese devices exhibit two-mode a uto-oscillations (Fig. 2(b)). T he frequency of the low-\nfrequency mode, excited at small currents, is smaller than the frequencies of propagating spin \nwaves in the 15 nm-thick waveguide. Correspondingly, the spatia l mapping of the oscillations 8\nby micro-focus BLS showed that this mode is localized in the na no-notch and does not emit \nspin waves into the waveguide (Fig. 3(c)). In contrast, the hig h-frequency mode, excited at \nlarger currents, was found to efficiently emit propagating spin waves (Fig. 3(d)). This emission \nwas found to be strongly unidirec tional, with the preferential direction controlled by the \ndirection of the static magne tic field. Additionally, it was sh own that the propagation length of \nemitted spin waves is enhanced by non-zero spin current injecte d over the entire length of the \nwaveguide, by up to a factor of three. \nThe system proposed in Ref. 62 combines all the advantages prov ided by SOT to locally \nexcite propagating spin waves, and to simultaneously enhance th eir propagation characteristics. \nThe achieved enhancement can be further increased by the materi al engineering and geometry \noptimization. Additionally, the proposed approach can be easily scaled to chains of SOT nano-\noscillators coupled via propaga ting spin waves, facilitating th e development of novel nanoscale \nsignal processing circuits such as logic and neuromorphic compu ting networks. We would like \nto note, however, that in spite of all these advantages, the pr oposed system also suffers from \nthe limitations associated with the nonlinear spin-wave scatter ing. The latter adversely affects \nthe oscillation characteristics of nano-notch devices at large driving currents, resulting in a \nstrong reduction of the intensity of emitted spin waves (Fig. 2 (b)), and does not allow one to \nachieve complete compensation of their propagation losses. \nIV. Control of nonlinear damp ing in SOT-driven devices \nThe effects of SOT on the magnetization are often approximated as a simple modification \nof the effective magnetic damping37. This simple picture neglects the effects of fluctuations \nalways present at finite temperatures. Analysis taking into acc ount this contribution reveals \nanother important effect of SOT – namely, it drives the spin sy stem out of thermal equilibrium, \nresulting in the enhancement of magnetic fluctuations, which ca n equivalently be described as \nexcitation of a large number of incoherent spin waves (magnons) spread over a broad interval 9\nof frequencies and wavelengths.39,67 The importance of this cont ribution has been recognized \nstarting with the first experimen ts on the interaction of spin currents with magnetization.49 Later \non, excitation of incoherent magnons by SOT and their propagati on received a significant \nattention as a mechanism for transmission of spin information i n magnetic insulators.68-72 \nHowever, the spectral characteri stics of magnons excited by SOT remained unknown for a long \ntime. This issue was addressed experimentally in Ref. 67. In th is work, by using the BLS \ntechnique, we were able to study the effects of SOT on the magn on distribution in a Py/Pt \nbilayer over a significant spectral range. \nFigure 3(a) shows the BLS spectra reflecting the spectral densi ty of magnons recorded with \nand without dc current in Pt. The spectrum obtained at I = 0 characterizes the magnons present \nin Py due to thermal fluctuations at room temperature. When the current is applied, the \npopulations of all magnon states s ignificantly increase due to the effects of SOT. Note that the \nenhancement of the magnon population is most significant at low f r e q u e n c i e s , a n d r a p i d l y \ndecreases with the increase of the frequency of magnons (Fig. 3 (b)), which can be associated \nwith the larger relaxation rate s of magnons with higher frequen cies. With the increase in the \nSOT strength, the dominant mode at the lowest frequency fmin is expected to transition to the \nauto-oscillation regime at its damping compensation point. Howe ver, the increase in the \namplitudes of all the other mode s leads to their nonlinear coup ling, which results in the energy \nflow from the dominant mode (see, e.g., results of micromagneti c simulations Fig. 5d in Ref. \n64). This process represents an onset of additional (nonlinear) damping, which prevents \ncomplete damping compensation by SOT. \nThe adverse effects of nonlinear damping can be reduced by supp ressing the amplitudes of \nparasitic incoherent modes. This approach was used to achieve S OT-driven coherent \nmagnetization oscillations by utilizing local injection of spin current.39,50 In this geometry, \nparasitic incoherent spin waves quickly escape from the localiz ed active area, resulting in 10\nreduced nonlinear damping of the dominant mode. This approach g enerally restricts the active \nregion to nanoscale dimensions, limiting the achievable dynamic al coherence and the \npossibilities for the magnonic device integration. Moreover, th is approach cannot be extended \nto the compensation of propagati on losses of spin waves by SOT over extended spatial areas. \nWe have recently proposed a new, more efficient approach to sup pression of nonlinear \ndamping, based on the direct control of the mode coupling mecha nisms.64 In this work, we \nshowed that the nonlinear spin wave coupling is predominantly d etermined by the ellipticity of \nmagnetization precession, which is controlled in thin magnetic films by the demagnetizing \neffects. We demonstrated that, by using magnetic films with sui tably tailored perpendicular \nmagnetic anisotropy (PMA), one can compensate the dipolar aniso tropy and achieve almost \ncircular precession, resulting in suppression of nonlinear damp ing. As a result, we were able to \nachieve complete damping compensation and excitation of coheren t magnetization auto-\noscillations by SOT, in a simple system with uniform spatially- extended injection of spin \ncurrent. \nFigure 4(a) shows the layout of the test device studied in this work. It consists of a 5 nm-\nthick magnetic disk fabricated on top of an 8 nm-thick Pt strip , which plays the role of a spin-\ncurrent injector. We emphasize, that it is well known from earl y studies39,49, that in this \ngeometry it is impossible to achieve complete damping compensat ion and excitation of \nmagnetization auto-oscillations in magnetic systems with in-pla ne anisotropy, due to the \nadverse effects of the nonlinea r damping. Indeed , no transition to auto-oscillations is observed \nwhen Permalloy is used as the magnetic disk material (Fig. 4(b) ). When the current I is applied \nthrough the Pt strip, the recorded BLS spectra show an enhancem ent of magnetic fluctuations \nin Py. The intensity of fluctuations gradually increases with i ncreasing I at currents smaller than \nthe critical value IC. However, at I > IC, the intensity of fluctuations saturates, while their \nspectral width signi ficantly increases. 11\nThe observed behaviors change dramatically, if the magnetic dis k is made from a Co/Ni \nbilayer, with the relative thic knesses of Co and Ni adjusted so that the effective PMA field Ha \ncompensates the saturation magnetization 4 M of the bilayer (Fig. 4(c )). The compensation of \nthe dipolar anisotropy by PMA res ults in nearly c ircular magnet ization precession trajectory. \nThis can be contrasted with the magnetization precession of the Py disk, which exhibits a large \nellipticity (see the insets in Fi gs. 4(b) and 4(c)). As seen fr om Fig. 4(c), at I < IC, the effects of \nSOT in CoNi disk are very similar to those in the Py disk. How ever, at I > IC, a narrow intense \nspectral peak emerges for CoNi, marking a transition to the aut o-oscillation regime. This result \nshows that the nonlinear damping that prevents the onset of aut o-oscillations in the Py disk is \nsuppressed in CoNi. \nThe experimental results describ ed above, together with the mic romagnetic simulations of \nthese systems, clearly demonstrated a route for overcoming the limitations imposed by the \nnonlinear damping, by utilizing PMA materials with tailored ani sotropy strength. This approach \nallows one to achieve complete compensation of the magnetic dam ping, and excitation of \ncoherent magnetization auto-oscillations by SOT, without confin ing the spin-current injection \nto a nanoscale area. This can enable the implementation of SOT- driven oscillators with spatially \nextended active area, capable of generating microwave signals w ith technologically relevant \npower levels and coherence, and t hus circumventing the challeng es of phase locking of a large \nnumber of oscillators with nano- scale dimensions. The proposed approach also provides a route \nfor the implementation of spatially extended amplification of c oherent propagating spin waves. \nV. Highly efficient SOT-driven s pin-wave emission in YIG films with PMA \nThe main challenges in the application of the approach describe d above to metallic \nmultilayers with PMA are associated with their large magnetic d amping, and the strong spatial \ninhomogeneity of magnetic properties typical for these systems. Recent progress in the \ndeposition of high-quality nanometer-thick films of YIG can all ow one to overcome these 12\nlimitations. In particular, it was shown in Ref. 73 that Bi dop ing of ultrathin YIG films can \nfacilitate a large PMA controllable by the epitaxial strain and growth-induced anisotropies, \nwhile preserving their low-dampi ng characteristics. These prope rties make Bi-doped YIG films \nuniquely suitable for SOT-based magnonic devices utilizing supp ression of nonlinear damping \nby the anisotropy compensation. \nFigure 5(a) shows the layout of the SOT devi ce based on an exte nded 20-nm thick film of \nBi-doped YIG (BiYIG) (Bi 1Y2Fe5O12) with PMA.65 To achieve suppression of the nonlinear \ninteractions, the strength of PMA is tuned to exactly compensat e the dipolar anisotropy. The \ndevice utilizes a simple large-scale injector of spin current, formed by a 6-nm thick Pt strip line \nwith the width of 1 m and the length of 4 m. Similar simple spin-current injectors were \nutilized in many prior experime nts on YIG films without PMA.68-72 We emphasize that it was \nwell established in those studies, that SOT-driven emission of coherent propagating SW cannot \nbe achieved in in-plane magnetized YIG, even at large driving c urrents exceeding IC. Instead, \nthe dominant mode saturates in the vicinity of the point of com plete damping compensation, \nsimilar to the Py disks on Pt (Fig. 4(b)).72 \nIn contrast, the devices based on BiYIG were found to exhibit a clear transition to coherent \nauto-oscillations at moderate current densities in the Pt strip . By using local detection of the \nauto-oscillation spectra with BLS (Fig. 5(b)), we confirmed the high spatial coherence of the \nauto-oscillations, whose freque ncy remains unchanged over the e ntire active area above the 1×4 \nµm Pt injector. In contrast to most of the previously demonstra ted SOT oscillators, the intensity \nof the auto-oscillation peak m onotonically increases with the i ncrease of the current strength \nover a broad range of currents (Fig. 5(c)), consistent with sup pression of nonlinear interactions \nthat typically result in the deg radation of the oscillation cha racteristics at large current densities \n(see, e.g., Fig. 2(b)). 13\nAnother consequence of the precise compensation of the dipolar anisotropy by PMA is the \nabsence of the nonlinear frequenc y shift of the auto-oscillatio ns with the increase of their \nintensity (Fig. 5(c)). The spectral coherence of auto-oscillati ons is known to be significantly \nenhanced when the nonlinear shift is small, as it hence reduces the coupling between the \noscillation amplitude and its phase.74 In addition, the nonlinear frequency shift, common for in-\nplane magnetized devices (see, e.g., Fig. 2(b)), is well-known to result in self-localization of \nauto-oscillation into a nonlinear spin-wave bullet.75,76 In the absence of the nonlinear shift, the \nself-localization is not expected to occur, enabling emission o f coherent spin waves into the \nsurrounding film. Indeed, spatiall y resolved BLS measurements ( Fig. 5(d)) confirm efficient \nspin-wave emission from the active device area. As seen from th ese data, the oscillations \nexcited due to SOT ar e not localized in the area above the Pt l ine, but significantly extend into \nthe surrounding BiYIG film. \nAnalysis performed in Ref. 65 showed that the wavelength of the emitted spin waves is \nabout 300 nm, as determined by the difference in the PMA streng th in the bare BiYIG film and \nin the BiYIG/Pt bilayer. The wavelength can be controlled by va rying the difference between \nthese anisotropies. Additionally, the frequency of auto-oscilla tions and the wavelength of the \nemitted spin waves can be controlled by the current, in devices where the strength of PMA is \nincreased beyond the point of exac t compensation of the dipolar anisotropy. These conditions \ncorrespond to the positive nonlinear frequency shift, which all ows one to increase the auto-\noscillation frequency and decrease the wavelength of emitted sp in waves by the increase of the \ncurrent in the Pt line. \nThe results described above clearly demonstrate that suppressio n of the adverse nonlinear \ndamping in materials with PMA allows one to overcome the most s ignificant limitations \nhindering the practical realizati on of advantages provided by S OT in magnonics. We believe \nthat these findings should spur f urther progress in this field. 14\nVI. Conclusions \nDownscaling of magnonic devices p resents a large number of new challenges, which must \nbe addressed before spin-wave t echnology becomes a competitive alternative to conventional \nCMOS-based microelectronics. Si multaneously, downscaling provid es novel opportunities \nunavailable in traditional macros copic-scale spin-wave devices. Spin-orbit torque is one of the \nstriking examples of physical mechanisms that become particular ly efficient at nanoscale. \nUtilization of such phenomena is necessary for the successful d evelopment of nano-magnonics, \nenabling the implementation of nanodevices that are more effici ent and more attractive for real-\nworld applications than their previously demonstrated macroscop ic counterparts. Naturally, \nnovel physical phenomena, which have not yet been fully underst ood, require intense research, \noften revealing new unexpected d ifficulties and challenges as w ell as new opportunities. This \nis clearly the case with SOT-driven excitation and amplificatio n of spin waves. The results \nobtained in early experiments in the “small-amplitude” regime c ould be easily interpreted by \nusing a simple interpretation of the effects of SOT in terms of the variation of the effective \ndamping. However, with the development of highly-efficient SOT systems, where complete \ndamping compensation can be achieved, new challenges have emerg ed, which can only be \novercome based on the deep insi ght into nonlinear dynamic proce sses in strongly -driven spin \nsystems. We believe that the recently proposed approach discuss ed in this article will finally \nenable efficient integration of s pin-orbitronics and magnonics, and will accelerate new \ndevelopments in the latter field. \n ACKNOWLEDGMENTS \nThis work was supported in part by the Deutsche Forschungsgemei nschaft (Project No. \n423113162), the NSF Grant No. ECCS-1804198, by the French ANR G rants 15-CE08-0030-\n01 (ISOLYIG) and the ANR-18-CE24-0021 (MAESTRO). The authors acknowledge the \nessential contribution of the PhD students that have contribute d to some of the studies presented 15\nin this article: Martin Collet, Boris Divinskiy, Michael Evelt, Diane Gouéré, Hugo Merbouche, \nand Lucile Soumah. \n \nData Availability \nThe data that support the findings of this study are available from the corresponding author \nupon reasonable request. \n \nReferences \n \n1 J.D. Adam, “Analog signal proc essing with microwave magnetics, ” Proc. IEEE 76, 159-170 \n(1988). \n2 W. S. Ishak, “Magnetostatic wav e technology: A review,” Proc. IEEE 76, 171-187 (1988). \n3 P. Kaboš and V. S. Stalmachov, “Magnetostatic Waves and Their Application,” Chapman & \nHall, New York, 1994. \n4 A.G. Gurevich and G.A. Melkov, “Ma gnetization Oscillations and Waves,” CRC, New \nYork, 1996. \n5 D. D. Stancil and A. Prabhakar, “ Spin Waves: Theory and Applic ations,” Springer, New \nYork, 2009. \n6 A. Khitun, M. Bao, and K. L. Wang , “Magnonic logic circuits,” J. Phys. D 43, 264005 \n(2010). \n7 S. Dutta, S.-C. Chang, N. Kani , D. E. Nikonov, S. Manipatruni, I. A. Young, and A. \nNaeemi, “Non-volatile Clocked Spi n Wave Interconnect for Beyond -CMOS Nanomagnet \nPipelines,” Sci. Rep. 5, 9861 (2015). \n8 G. Csaba, A. Papp, and W. Porod, “Perspectives of using spin w aves for computing and \nsignal processing,” Phys. Lett. A 381, 1471 (2017). \n9 S. Neusser, D. Grundler, “Magnonics: Spin waves on the nanosca le,” Adv. Mater. 21, 2927–\n2932 (2009). \n10 V. V. Kruglyak, S. O. Demokritov, and D. Grundler, “Magnonics, “ J. Phys. D 43, 264001 \n(2010). \n11 B. Lenk, H. Ulrichs, F. Garbs, M. Münzenberg, “The building bl ocks of magnonics,” Phys. \nRep. 507, 107–136 (2011). 16\n12 A. Chumak, V. Vasyuchka, A. Serga, B. Hillebrands, “Magnon spi ntronics,” Nat. Phys. 11, \n453–461 (2015). \n13 B. A. Kalinikos, “Excitation of pr opagating spin waves in ferr omagnetic films,” IEE Proc. \nH. 127, 4–10 (1980). \n14 V. Vlaminck and M. Bailleul, “Spin-wave transduction at the su bmicrometer scale: \nExperiment and modeling,” Phys. Rev. B 81, 014425 (2010). \n15 V. E. Demidov, M. P. Kostylev, K. Rott, J. Münchenberger, G. R eiss, and S. O. \nDemokritov, “Excitation of short- wavelength spin waves in magno nic waveguides,” Appl. \nPhys. Lett. 99, 082507 (2011). \n16 H. Yu, O. d’ Allivy Kelly, V. Cro s, R. Bernard, P. Bortolotti, A. Anane, F. Brandl, F. \nHeimbach, and D. Grundler, “Approa ching soft X-ray wavelengths in nanomagnet-based \nmicrowave technology,” Nat. Commun. 7, 11255 (2016). \n17 C. Liu, J. Chen, T. Liu, F. H eimbach, H. Yu, Y. Xiao, J. Hu, M . Liu, H. Chang, T. \nStueckler, S. Tu, Y. Zhang, Y. Zhang, P. Gao, Z. Liao, D. Yu, K . Xia, N. Lei, W. Zhao, \nand M. Wu, “Long-distance propaga tion of short-wavelength spin waves,” Nat. Commun. \n9, 738 (2018). \n18 A. Papp, G. Csaba, H. Dey, M. Madami, W. Porod, and G. Carlott i, “Waveguides as \nsources of short-wavelength spin w aves for low-energy ICT appli cations,” Eur. Phys. J. B \n91, 107 (2018). \n19 P. Pirro, T. Brächer, A. V. Chu mak, B. Lägel, C. Dubs, O. Surz henko, P. Görnert, B. \nLeven, and B. Hillebrands, “Spin-wave excitation and propagatio n in microstructured \nwaveguides of yttrium iron garnet /Pt bilayers,” Appl. Phys. Let t. 104, 012402 (2014). \n20 M. B. Jungfleisch, W. Zhang, W. Jiang, H. Chang, J. Sklenar, S . M. Wu, J. E. Pearson, A. \nBhattacharya, J. B. Ketterson, M . Wu, and A. Hoffmann, “Spin wa ves in micro-structured \nyttrium iron garnet nanometer-thick films,” J. Appl. Phys. 117, 17D128 (2015). \n21 M. Collet, O. Gladii, M. Eve lt, V. Bessonov, L. Soumah, P. Bor tolotti, S. O. Demokritov, \nY. Henry, V. Cros, M. Bailleul, V. E. Demidov, and A. Anane, “S pin-wave propagation in \nultra-thin YIG based wave guides,” Appl. Phys. Lett. 110, 092408 (2017). \n22 J. C. Slonczewski, “Current-drive n excitation of magnetic mult ilayers,” J. Magn. Magn. \nMater. 159, L1–L7 (1996). \n23 L. Berger, “Emission of spin waves by a magnetic multilayer tr aversed by a current,” Phys. \nRev. B 54, 9353–9358 (1996). 17\n24 J. A. Katine, F. J. Albert, R . A. Buhrman, E. B. Myers, and D. C. Ralph, “Current-Driven \nMagnetization Reversal and Spin-Wa ve Excitations in Co/Cu/Co Pi llars,” Phys. Rev. Lett. \n84, 3149-3152 (2000). \n25 M. Tsoi, A. G. M. Jansen, J. Bass, W.-C. Chiang, V. Tsoi, and P. Wyder, “Generation and \ndetection of phase-coherent curr ent-driven magnons in magnetic multilayers,” Nature 406, \n46-48 (2000). \n26 V. E. Demidov, S. Urazhdin and S . O. Demokritov, “Direct obser vation and mapping of \nspin waves emitted by spin-torque nano-oscillators,” Nature Mat er. 9, 984-988 (2010). \n27 M. Madami, S. Bonetti, G. Cons olo, S. Tacchi, G. Carlotti, G. Gubbiotti, F. B. Mancoff, M. \nA. Yar, J. Åkerman, “Direct obser vation of a propagating spin w ave induced by spin-\ntransfer torque,” Nat. Nanotech. 6, 635–638 (2011). \n28 H. Ulrichs, V. E. Demidov, S. O . Demokritov, and S. Urazhdin, “Spin-torque nano-emitters \nfor magnonic applications,” Appl. Phys. Lett. 100, 162406 (2012). \n29 S. Urazhdin, V. E. Demidov, H. Ulr ichs, T. Kendziorczyk, T. Ku hn, J. Leuthold, G. Wilde \nand S. O. Demokritov, “Nanomagnoni c devices based on the spin-t ransfer torque,” Nat. \nNanotech. 9, 509-513 (2014). \n30 V. E. Demidov, S. Urazhdin, R. Li u, B. Divinskiy, A. Telegin, and S. O. Demokritov, \n“Excitation of coherent propaga ting spin waves by pure spin cur rents,” Nat. Commun. 7, \n10446 (2016). \n31 B. Divinskiy, V. E. Demidov, S. O. Demokritov, A. B. Rinkevich , and S. Urazhdin, “Route \ntoward high-speed nano-magnonics provided by pure spin currents ,” Appl. Phys. Lett. 109, \n252401 (2016). \n32 A. Houshang, R. Khymyn, H. Fulara , A. Gangwar, M. Haidar, S. R . Etesami, R. Ferreira, P. \nP. Freitas, M. Dvornik, R. K. Dum as, J. Åkerman, “Spin transfer torque driven higher-\norder propagating spin waves in nano-contact magnetic tunnel ju nctions,” Nat. Commun. \n9, 4374 (2018). \n33 A. Hoffmann, “Spin Hall Effects in Metals,” IEEE Trans. Magn. 49, 5172–5193 (2013). \n34 J. Sinova, S. O. Valenzuela , J. Wunderlich, C. H. Back, and T. Jungwirth, “Spin Hall \neffects,” Rev. Mod. Phys. 87, 1213 (2015). \n35 A. Soumyanarayanan, N. Reyre n, A. Fert, and C. Panagopoulos, “ Emergent phenomena \ninduced by spin–orbit coupling at surfaces and interfaces,” Nat ure 539, 509–517 (2016). \n36 A. Manchon, J. Železný, I.M. M iron, T. Jungwirth, J. Sinova, A. Thiaville, K. Garello, and \nP. Gambardella, “Curre nt-induced spin-orbit torques in ferromag netic and \nantiferromagnetic systems,” Rev. Mod. Phys. 91, 035004 (2019). 18\n37 K. Ando, S. Takahashi, K. Harii, K. Sasage, J. Ieda, S. Maekaw a, and E. Saitoh, “Electric \nManipulation of Spin Relaxation U sing the Spin Hall Effect,” Ph ys. Rev. Lett. 101, 036601 \n(2008). \n38 Y. Kajiwara, K. Harii, S. Takah ashi, J. Ohe, K. Uchida, M. Miz uguchi, H. Umezawa, H. \nKawai, K. Ando, K. Takanashi, S. M aekawa, and E. Saitoh, “Trans mission of electrical \nsignals by spin-wave interconvers ion in a magnetic insulator,” Nature 464, 262–266 \n(2010). \n39 V. E. Demidov, S. Urazhdin, G. de Loubens, O. Klein, V. Cros, A. Anane, and S. O. \nDemokritov, “Magnetization osci llations and waves driven by pur e spin currents,” Phys. \nRep. 673, 1-31 (2017). \n40 Z. Wang, Y. Sun, M. Wu, V. Tiberke vich, and A. Slavin, “Contro l of SpinWaves in a Thin \nFilm Ferromagnetic Insulator th rough Interfacial Spin Scatterin g,” Phys. Rev. Lett. 107, \n146602 (2011). \n41 E. Padron-Hernandez, A. Azevedo, and S. M. Rezende, “Amplifica tion of spin waves in \nyttrium iron garnet films through t he spin Hall effect,” Appl. Phys. Lett. 99, 192511 \n(2011). \n42 L.J. Cornelissen, J. Liu, B.J. van Wees, and R.A. Duine, “S pin-Current-Controlled \nModulation of the Magnon Spin Conduc tance in a Three-Terminal M agnon Transistor,” \nPhys. Rev. Lett. 120, 097702 (2018). \n43 T. Wimmer, M. Althammer, L. Lie nsberger, N. Vlietstra, S. Gepr ägs, M. Weiler, R. Gross, \nand H. Huebl, “Spin Transport in a Magnetic Insulator with Zero Effective Damping,” \nPhys. Rev. Lett. 123, 257201 (2019). \n44 V. E. Demidov, S. Urazhdin, A. B . Rinkevich, G. Reiss, and S. O. Demokritov, “Spin Hall \ncontrolled magnonic microwaveguides,” Appl. Phys. Lett. 104, 152402 (2014). \n45 K. An, D. R. Birt, C.-F. Pai , K. Olsson, D. C. Ralph, R. A. Bu hrman, and X. Li, “Control of \npropagating spin waves via spin transfer torque in a metallic b ilayer waveguide,” Phys. \nRev. B 89, 140405(R) (2014). \n46 M. Evelt, V. E. Demidov, V. B essonov, S. O. Demokritov, J. L. Prieto, M. Muñoz, J. Ben \nYoussef, V. V. Naletov, G. de L oubens, O. Klein, M. Collet, K. Garcia-Hernandez, P. \nBortolotti, V. Cros and A. Anane , “High-efficiency control of s pin-wave propagation in \nultra-thin yttrium iron garnet by the spin-orbit torque,” Appl. Phys. Lett. 108, 172406 \n(2016). 19\n47 O. Gladii, M. Collet, K. Garcia- Hernandez, C. Cheng, S. Xavier , P. Bortolotti, V. Cros, Y. \nHenry, J.-V. Kim, A. Anane, and M. Bailleul, “Spin wave amplifi cation using the spin Hall \neffect in permalloy/platinum bilayers,” Appl. Phys. Lett. 108, 202407 (2016). \n48 A. Navabi, Y. Liu, P. Upadhyaya, K . Murata, F. Ebrahimi, G. Yu , Bo Ma, Y. Rao, M. \nYazdani, M. Montazeri, Lei Pan, I. N. Krivorotov, I. Barsukov, Q. Yang, P. K. Amiri, Y. \nTserkovnyak, and K. L. Wang, “Contro l of Spin-Wave Damping in Y IG Using Spin \nCurrents from Topological Insul ators,” Phys. Rev. Applied 11, 034046 (2019). \n49 V. E. Demidov, S. Urazhdin, E. R . J. Edwards, M. D. Stiles, R. D. McMichael, and S. O. \nDemokritov, “Control of magnetic fluctuations by spin current,” Phys. Rev. Lett. 107, \n107204 (2011). \n50 V. E. Demidov, S. Urazhdin, H. Ulri chs, V. Tiberkevich, A. Sla vin, D. Baither, G. Schmitz, \nand S. O. Demokritov, “Magnetic na no-oscillator driven by pure spin current,” Nature \nMater. 11, 1028-1031 (2012). \n51 L. Liu, C.-F. Pai, D. C. Ralph, R . A. Buhrman, “Magnetic Oscil lations Driven by the Spin \nHall Effect in 3-Terminal Magne tic Tunnel Junction Devices,” Ph ys. Rev. Lett. 109, \n186602 (2012). \n52 V. E. Demidov, S. Urazhdin, A . Zholud, A. V. Sadovnikov, and S . O. Demokritov, \n“Nanoconstriction-based spin-Hal l nano-oscillator,” Appl. Phys. Lett. 105, 172410 (2014). \n53 Z. Duan, A. Smith, L. Yang, B. Youngblood, J. Lindner, V. E. D emidov, S. O. Demokritov, \nand I. N. Krivorotov, “Nanowire sp in torque oscillator driven b y spin orbit torques,” Nat. \nCommun. 5, 5616 (2014). \n54 M. Collet, X. de Milly, O. d’Allivy Kelly, V.V. Naletov, R. Be rnard, P. Bortolotti, J. Ben \nYoussef, V. E. Demidov, S. O. Dem okritov, J. L. Prieto, M. Muno z, V. Cros, A. Anane, G. \nde Loubens, and O. Klein, “Genera tion of coherent spin-wave mod es in yttrium iron garnet \nmicrodiscs by spin–orbit torque,” Nat. Commun. 7, 10377 (2016). \n55 T. Chen, R. K. Dumas, A. Ekl und, P. K. Muduli, A. Houshang, A. A. Awad, P. Dürrenfeld, \nB. G. Malm, A. Rusu, J. Åkerman, “ Spin-torque and spin-Hall nan o-oscillators,” Proc. \nIEEE 104, 1919–1945 (2016). \n56 P. Dürrenfeld, A. A. Awad, A. H oushang, R. K. Dumas, and J. Åk erman, “A 20 nm spin \nHall nano-oscillator,” Nanoscale 9, 1285–1291 (2017). \n57 B. Divinskiy, V. E. Demidov, A. Kozhanov, A. B. Rinkevich, S. O. Demokritov, and S. \nUrazhdin, “Nanoconstriction spin-Ha ll oscillator with perpendic ular magnetic anisotropy,” \nAppl. Phys. Lett. 111, 032405 (2017). 20\n58 A. A. Awad, P. Dürrenf eld, A. Houshang, M. Dvornik, E. Iacocca , R. K. Dumas, and J. \nÅkerman, “Long-range mutual sync hronization of spin Hall nano-o scillators,” Nat. Phys. \n13, 292 (2017). \n59 B. Divinskiy, V. E. Demidov, S. Urazhdin, R. Freeman, A. B. Ri nkevich, and S. O. \nDemokritov, “Controllable exc itation of quasi-l inear and bullet modes in a spin-Hall nano-\noscillator,” Appl. Phys. Lett. 114, 042403 (2019). \n60 N. Sato, K. Schultheiss, L. Kör ber, N. Puwenberg, T. Mühl, A. A. Awad, S.S.P.K. \nArekapudi, O. Hellwig, J. Fassb ender, and H. Schultheiss, “Doma in Wall Based Spin-Hall \nNano-Oscillators,” Phys. Rev. Lett. 123, 057204 (2019). \n61M. Zahedinejad, A. A. Awad, S. Muralidhar, R. Khymyn, H. Fulara , H. Mazraati, M. \nDvornik, and J. Åkerman, “Two-dime nsional mutually synchronized spin Hall nano-\noscillator arrays for neuromor phic computing,” Nat. Nanotech. 15, 47-52 (2020). \n62 B. Divinskiy, V. E. Demidov, S. Urazhdin, R. Freeman, A. B. Ri nkevich, and S. O. \nDemokritov, “Excitation and Amp lification of Spin Waves by Spin –Orbit Torque,” Adv. \nMater. 30, 1802837 (2018). \n63 H. Fulara, M. Zahedinejad, R . Khymyn, A. A. Awad, S. Muralidha r, M. Dvornik, and J. \nÅkerman, “Spin-orbit torque–driven propagating spin waves,” Sci . Adv. 5, eaax8467 \n(2019). \n64 B. Divinskiy, S. Urazhdin, S. O . Demokritov, and V. E. Demidov , “Controlled nonlinear \nmagnetic damping in spin-Hall nano-devices,” Nat. Commun. 10, 5211 (2019). \n65 M. Evelt, L. Soumah, A. B. Ri nkevich, S. O. Demokritov, A. Ana ne, V. Cros, J. Ben \nYoussef, G. de Loubens, O. Klei n, P. Bortolotti, and V. E. Demi dov, “Emission of \ncoherent propagating magnons by ins ulator-based spin-orbit-torq ue oscillators,” Phys. Rev. \nAppl. 10, 041002 (2018). \n66 V. E. Demidov and S. O. Demokritov, “Magnonic waveguides studi ed by micro-focus \nBrillouin light scattering,” IEEE Trans. Mag. 51, 0800215 (2015). \n67 V. E. Demidov, S. Urazhdin, B. Divinskiy, V. D. Bessonov, A. B . Rinkevich, V.V. \nUstinov, and S. O. Demokritov, “C hemical potential of quasi-equ ilibrium magnon gas \ndriven by pure spin current,” Nat. Commun. 8, 1579 (2017). \n68 L. J. Cornelissen, J. Liu, R. A. Duine, J. Ben Youssef, and B. J. van Wees, “Long-distance \ntransport of magnon spin informa tion in a magnetic insulator at room temperature,” Nat. \nPhys. 11, 1022-1026 (2015). 21\n69 S. T. B. Goennenwein, R. Schlitz, M. Pernpeintner, K. Ganzhorn , M. Althammer, R. Gross, \nand H. Huebl, “Non-local magnetore sistance in YIG/Pt nanostruct ures,” Appl. Phys. Lett. \n107, 172405 (2015). \n70 L. J. Cornelissen, K. J. H. Peter s, G. E. W. Bauer, R. A. Duin e, and B. J. van Wees, \n“Magnon spin transport drive n by the magnon chemical potential in a magnetic insulator,” \nPhys. Rev. B 94, 014412 (2016). \n71 D. Wesenberg, T. Liu, D. Balzar, M. Wu, and B. L. Zink, “Long- distance spin transport in a \ndisordered magnetic ins ulator,” Nat. Phys. 13, 987 (2017). \n72 N. Thiery, A. Draveny, V. V. Nal etov, L. Vila, J. P. Attané, C . Beigné, G. de Loubens, M. \nViret, N. Beaulieu, J. Ben You ssef, V. E. Demidov, S. O. Demokr itov, A. N. Slavin, V. S. \nTiberkevich, A. Anane, P. Borto lotti, V. Cros, and O. Klein, “ Nonlinear spin conductance \nof yttrium iron garnet thin films driven by large spin-orbit to rque,” Phys. Rev. B 97, \n060409 (2018). \n73 L. Soumah, N. Beaulieu, L. Qassy m, C. Carrétéro, E. Jacquet, R . Lebourgeois, J. Ben \nYoussef, P. Bortolotti, V. Cros , and A. Anane, “Ultra-low dampi ng insulating magnetic \nthin films get perpendicular,” Nat. Commun. 9, 3355 (2018). \n74 A. Slavin and V. Tiberkevich, “No nlinear Auto-Osc illator Theor y of Microwave \nGeneration by Spin-Polarized C urrent,” IEEE Trans. Magn. 45, 1875-1918 (2009). \n75 A. N. Slavin and V. S. Tiberkevich, “Spin Wave Mode Excited by Spin-Polarized Current \nin a Magnetic Nanocontact is a S tanding Self-Localized Wave Bul let,” Phys. Rev. Lett. 95, \n237201 (2005). \n76 M. Dvornik, A. A. Awad, and J. Åkerman, “Origin of Magnetizati on Auto-Oscillations in \nConstriction-Based Spin Hall Na no-Oscillators,” Phys. Rev. Appl . 9, 014017 (2018). \n 22\n \n \n \n \n \n \n \n \nFig. 1. (a) Schematic of the SOT device based on the Pt/YIG bil ayer waveguide, and the \nexperimental setup. (b) Measured spatial map of the BLS intensi ty, which is proportional \nto the local spin-wave intensity. (c) Dependences of the spin-w ave intensity on the \npropagation coordinate, recorded at different dc currents in th e Pt layer, as labeled. \nSymbols – experimental data, curves – exponential fits. (d) Cur rent dependences of the \ndecay length (point-up triangles) and of its inverse value (poi nt-down triangles). IC marks \nthe critical current, at which S OT is expected to completely co mpensate the natural \ndamping. Reproduced from M. Evelt, V. E. Demidov, V. Bessonov, S. O. Demokritov, J. \nL. Prieto, M. Muñoz, J. Ben Youssef, V. V. Naletov, G. de Loube ns, O. Klein, M. Collet, \nK. Garcia-Hernandez, P. Bortolotti, V. Cros and A. Anane, Appl. Phys. Lett. 108, 172406 \n(2016), with the permission of AIP Publishing. \n \n23\n \n \n \n \n \n \n \n \nFig. 2. (a) Schematic of the SOT device base d on the notched Pt/Py bilayer nano-waveguide, \nand the experimental setup. (b) Normalized color-coded map of the BLS intensity in the \nfrequency-current coordinates. (c) and (d) Color-coded spatial maps of the BLS intensity, \nmeasured at the frequencies of t he localized and of the radiati ng modes, as labeled. Dashed \nlines on the maps show the outlines of the waveguide and of the nano-notch. Reproduced \nwith permission from B. Divinskiy, V. E. Demidov, S. Urazhdin, R. Freeman, A. B. \nRinkevich, and S. O. Demokritov, Adv. Mater. 30, 1802837 (2018). Copyright 2018 John \nWiley and Sons. \n \n24\n \n \n \n \n \n \n \n \nFig. 3. Effects of SOT on spectral magnon distribution for an S OT device based on crossed Pt \nand Py strips. (a) BLS spectra, reflecting the spectral density of magnons, recorded at I=0 and \n20 mA, as labeled. Vertical dashe d line marks the frequency of the lowest-energy magnon \nstate. (b) Enhancement of the magnon population at I=20 mA. Reproduced from V. E. \nDemidov, S. Urazhdin, B. Divinskiy, V. D. Bessonov, A. B. Rinke vich, V.V. Ustinov, and S. \nO. Demokritov, Nat. Commun. 8, 1579 (2017); licensed under a Creative Commons \nAttribution (CC BY) license. \n \n25\n \n \n \n \n \n \n \n \n \n \nFig. 4. (a) Layout of the test SOT devices based on Py and CoNi disks on th e Pt strip. (b) and \n(c) BLS spectra of magnetic oscillations vs current for Py and CoNi disks, respectively. IC \nmarks the critical current, at which SOT is expected to complet ely compensate the natural \nlinear magnetic damping. Insets illustrate the ellipticities o f magnetization precession in Py \nand CoNi, respectively. Reproduced from B. Divinskiy, S. Urazhd in, S. O. Demokritov, and \nV. E. Demidov, Nat. Commun. 10, 5211 (2019); licensed under a Creative Commons \nAttribution (CC BY) license. \n \n26\n\n\n\n\n\n\n\n\nFig. 5. (a) Schematic of the SOT devices utilizing Pt strip on extended BiY IG film . (b) \nRepresentative BLS spectrum of m agnetization auto-oscillations recorded at I=2.5 mA. Note \nthat the width of the measured spectral peak is determined by t he limited frequency resolution \nof BLS. (c) Color-coded BLS inte nsity in the current-frequency coordinates. (d) Color-coded \nspatial map of the BLS intensity recorded by rastering the prob ing laser spot over 20 m by 7 \nm area. Dashed lines show the c ontours of the Pt line. Reproduc ed with permission from M. \nEvelt, L. Soumah, A. B. Rinkevich , S. O. Demokritov, A. Anane, V. Cros, J. Ben Youssef, G. \nde Loubens, O. Klein, P. Bortolo tti, and V. E. Demidov, Phys. R ev. Appl. 10, 041002 (2018). \nCopyright 2018 by the American Physical Society. \n\n\n \n" }, { "title": "2304.00528v2.Interedge_spin_resonance_in_the_Kitaev_quantum_spin_liquid.pdf", "content": "Interedge spin resonance in the Kitaev quantum spin liquid\nTakahiro Misawa,1Joji Nasu,2and Yukitoshi Motome3\n1Beijing Academy of Quantum Information Sciences, Haidian District, Beijing 100193, China\n2Department of Physics, Tohoku University, Sendai 980-8578, Japan\n3Department of Applied Physics, University of Tokyo, Tokyo 113-8656, Japan\n(Dated: September 11, 2023)\nThe Kitaev model offers a platform for quantum spin liquids (QSLs) with fractional excitations,\nitinerant Majorana fermions and localized fluxes. Since these fractional excitations could be utilized\nfor quantum computing, how to create, observe, and control them through the spin degree of freedom\nis a central issue. Here, we study dynamical spin transport in a wide range of frequency for the\nKitaev-Heisenberg model, by applying an AC magnetic field to an edge of the system. We find that,\nin the Kitaev QSL phase, spin polarizations at the other edge are resonantly induced in a specific\nspin component, even though the static spin correlations are vanishingly small. This interedge spin\nresonance appears around the input frequency over the broad frequency range. Comparing with the\ndynamical spin correlations, we clarify that the resonance is governed by the itinerant Majorana\nfermions with a broad continuum excitation spectrum, which can propagate over long distances,\nalthough it vanishes for the pure Kitaev model because of accidental degeneracy and requires weak\nHeisenberg interactions. We also find that the spin polarizations in the other spin components are\nweakly induced at an almost constant frequency close to the excitation gap of the localized fluxes,\nirrespective of the input frequency. These results demonstrate that the dynamical spin transport is\na powerful probe of the fractional excitations in the Kitaev QSL. Possible experimental realization\nof the interedge spin resonance is discussed.\nI. INTRODUCTION\nExotic quasiparticles emerging in solids have attracted\nmuch interest from both fundamental physics and in-\ndustry applications. A prominent example is Majorana\nparticles — charge-neutral spin-1 /2 particles that are\ntheir own antiparticles1,2. While they usually behave\nas fermions, in some two-dimensional cases they can\nbe regarded as anyons that obey neither Fermi-Dirac\nnor Bose-Einstein statistics. Such Majorana particles\nhave been intensively studied for applications to quan-\ntum computing by using the anyonic nature3,4.\nThe Kitaev model on a honeycomb lattice offers an\nideal platform for realizing the Majorana particles5. It is\nan exactly solvable model whose ground state is a quan-\ntum spin liquid (QSL). The Kitaev QSL hosts two types\nof emergent quasiparticles from the fractionalization of\nspins: itinerant Majorana fermions and localized fluxes.\nThese quasiparticles turn into Abelian anyons when the\ninteractions between spins are largely anisotropic, or non-\nAbelian anyons when an external magnetic field is ap-\nplied in the nearly isotropic cases3,5. The Kitaev model\ncould be realized in Mott insulators with the strong spin-\norbit coupling6, such as Na 2IrO37andα-RuCl 38. De-\ntailed comparisons between experimental results and the-\noretical calculations have revealed fingerprints of the Ma-\njorana particles in such candidate materials; for a re-\nview, see Ref. 9. Among them, the discovery of the\nhalf-quantized thermal Hall effect in α-RuCl 3was ground\nbreaking, providing direct evidence for the Majorana par-\nticles10–12, while it is still under debate13–15.\nSince the Majorana fermions and the fluxes in the Ki-\ntaev QSL are generated by the fractionalization of spins,\nthey are quantum entangled and inherently nonlocal.Indeed, the spatial correlations between the Majorana\nfermions are long-range with power-law decay16,17, al-\nthough the spin correlations are short-ranged and van-\nish beyond nearest-neighbor sites18. Furthermore, in the\npresence of defects or edges, the spin correlations can\nbe long-range due to low-energy excitations around the\ndefects or edges16,17,19–21.\nBy exploiting such nonlocal nature, it was recently\nshown that the itinerant Majorana fermions can con-\ntribute to long-range spin transport from an edge of the\nsystem22,23. The previous study has focused only on\nlow-energy properties, such as the velocity of the spin\npropagation determined by the slope of the gapless Majo-\nrana dispersion. However, the spin dynamics in the wider\nrange of frequency is expected to offer more important in-\nsights into the two types of fractional quasiparticles with\ndistinct excitation spectra. Such comprehensive study\nof nonlocal spin dynamics would also be a crucial step\ntoward quantum computing, by elucidating how to con-\ntrol and probe the fractional quasiparticles via the spin\ndegree of freedom.\nIn this paper, in order to deepen the understanding\nof the relationship between the fractional quasiparticles\nand the spin degree of freedom, we study nonlocal spin\ndynamics in the Kitaev QSL in the wide frequency range.\nApplying a local AC magnetic field to one edge of the sys-\ntem, we investigate how the spin excitations are excited\nand propagate to the other edge. At the edges of the Ki-\ntaev model, it is known that local magnetic fields excite\nthe fluxes accompanied by gapless Majorana excitations,\ncalled the Majorana zero modes5,16. In the present study,\nwe introduce not static but time-dependent local mag-\nnetic fields at one edge and investigate how the excited\nspin polarizations propagate through the system. FromarXiv:2304.00528v2 [cond-mat.str-el] 8 Sep 20232\n(a) armchair (b) zigzag\nxy\nzhin(t)\nSiout(t)hin(t) Siout(t)\nFIG. 1. Schematic pictures of the Kitaev-Heisenberg model in\nEq. (1) with (a) the armchair edges and (b) the zigzag edges.\nIn (a) [(b)], we set the periodic and open (open and periodic)\nboundary conditions in the horizontal and vertical directions,\nrespectively. The blue, green, and red bonds represent the x,\ny, and zbonds for the Kitaev interaction, respectively. We\napply an AC magnetic field hin(t) in the [111] direction in\nspin space at one site on the edge shown by the yellow circles,\nand observe spin polarization at a site on the opposite edge\nshown by the orange circles; see Eqs. (3), (4), and (6).\nthe comprehensive analysis of the spin-component depen-\ndence and the comparison with the results for magnet-\nically ordered phases, we show that the dynamical spin\ntransport is a good probe for both itinerant Majorana\nfermions and localized fluxes. Our results give an insight\non the way of creating and controlling of the fractional\nexcitations via the spin degree of freedom.\nThe organization of this paper is as follows. In Sec. II,\nwe introduce the model and the setup used in this study,\nwith the details of real-time evolution and the definitions\nof static and dynamical spin correlations. In Sec. III A,\nwe show the phase diagram and the static spin correla-\ntions in our model with edges. In Sec. III B, we show\nhow an AC local magnetic field at the edge induces the\nspin polarization at the opposite edge of the system in\nthe ferromagnetic phase, the Kitaev QSL, and the stripy\nphase. In Sec. III C, we analyze the results in compari-\nson with the dynamical spin correlations between edges,\nand discuss the origin of the interedge dynamical spin\ntransport. Section IV is devoted to a summary.\nII. MODEL AND METHOD\nIn this paper, we employ the Kitaev-Heisenberg model,\nwhose Hamiltonian is given by\nˆHKH=KX\nνX\n⟨i,j⟩νˆSν\niˆSν\nj+JX\n⟨i,j⟩ˆSi·ˆSj, (1)\nwhere ˆSν\nidenotes the νth component of the spin-1/2\noperator at ith site: ˆSi= (ˆSx\ni,ˆSy\ni,ˆSz\ni). The first\nterm represents the bond-dependent Ising-type interac-\ntion, called the Kitaev interaction, where ⟨i, j⟩νrep-resents the nearest-neighbor ν(=x, y, z ) bonds on the\nhoneycomb lattice, and the second term represents the\nspin-isotropic Heisenberg interaction for all the nearest-\nneighbor bonds; see Fig. 1. Following the previous stud-\nies7,24, we parametrize the two coupling constants as\n(K, J) =\u0012\nsinα,1\n2cosα\u0013\n. (2)\nNote that the amplitudes of interactions are halved com-\npared to the previous ones so that |K|= 1 in the pure\nKitaev cases with α=π/2 and 3 π/2. In the follow-\ning, we focus on the range of π≤α≤7π/4 where\nthe Kitaev interaction is ferromagnetic. In this region,\nthe bulk system with the periodic boundary conditions\nshows three phases in the ground state24: the ferro-\nmagnetic phase for π≤α≲1.40π, the Kitaev QSL\nphase for 1 .40π≲α≲1.58π, and the stripy phase for\n1.58π≲α≲1.81π.\nTo study spin correlations and dynamics on the edges,\nwe consider the model in Eq. (1) on a strip with the\nopen boundary condition in one direction and the pe-\nriodic boundary condition in the other. There are two\ntypes of such strips. One has the so-called armchair type\nedges on the open boundaries, and the other has the so-\ncalled zigzag edges. Figure 1 displays these two types for\n24-site clusters used in the following calculations. For\nboth clusters, we examine how a time-dependent local\nmagnetic field on one edge induces spin polarization at\nthe other edge. Specifically, we apply an AC magnetic\nfield in the [111] direction in spin space at iinth site on\nthe edge (shown by the yellow circle in Fig. 1) as\nˆH(t) =ˆHKH+hin(t)·ˆSiin, (3)\nwith\nhin(t) =heccos (Ω t), (4)\nwhere his the amplitude of the AC field, ec=\n(1,1,1)/√\n3, and Ω = 2 π/T is the frequency of the AC\nfield ( Trepresents the oscillation period). We take the\nmagnetic field along the [111] direction since it coupled\nto all spin components. For this Hamiltonian, we solve\nthe time-dependent Schr¨ odinger equation given by\nid|Φ(t)⟩\ndt=ˆH(t)|Φ(t)⟩, (5)\nstarting from the initial condition of |Φ(t= 0)⟩=|ΦGS⟩,\nwhere |ΦGS⟩is the normalized ground state of ˆHKH. The\nspin polarization on the opposite edge is calculated as\nSν\niout(t) =⟨Φ(t)|ˆSν\niout|Φ(t)⟩, (6)\nwhere ioutdenotes the site on the other edge directly\nopposite to the iinth site (shown by the orange circle in\nFig. 1). In the following calculations, we take h= 0.05 in\nEq. (4) and solve Eq. (5) by using HΦ25; we discretize the\ntime with ∆ t= 0.05, which is small enough to preserve\nthe unitarity of real-time evolution.3\nIn addition to the real-time dynamics of the spin polar-\nization, we calculate the static and dynamical spin corre-\nlations between the two edges for the ground state |ΦGS⟩,\nwhich are defined by\nCνν\nedge=⟨ΦGS|ˆSν\niinˆSν\niout|ΦGS⟩, (7)\nand\nCνν\nedge(ω) =1\n2πZ∞\n−∞⟨ΦGS|δˆSν\niin(t)δˆSν\niout|ΦGS⟩eiωtdt,\n(8)\nrespectively, where δˆSν\ni=ˆSν\ni− ⟨ΦGS|ˆSν\ni|ΦGS⟩and\nδˆSν\niin(t) =eiˆHKHtδˆSν\niine−iˆHKHt. Note that in Eq. (8) we\nconsider correlations between the deviations from the ex-\npectation values for the ground state to subtract the elas-\ntic components in the presence of magnetic ordering. In\nthe actual calculations of Eq. (8), we employ the follow-\ning formula in the spectral representation:\nCνν\nedge(ω)\n=−1\n4πImh\n⟨Ψ+|(EGS−ˆHKH+ω+iη)−1|Ψ+⟩\n−⟨Ψ−|(EGS−ˆHKH+ω+iη)−1|Ψ−⟩i\n,\n(9)\nwhere |Ψ±⟩= (δˆSν\niin±δˆSν\niout)|ΦGS⟩,EGSis the ground-\nstate energy, and ηis an infinitesimal positive constant;\nwe take η= 0.05 in the following calculations. We calcu-\nlate Eq. (9) using the continued-fraction expansion based\non the Lanczos method.\nIn the calculations of Eqs. (7) and (9), we apply a\nweak static magnetic field to all the spins at the edge\non the iinth side with hs= 0.005ecto lift the ground-\nstate degeneracy in the ferromagnetic Heisenberg model\nwith α=πand the pure Kitaev model with α= 3π/2\n(see Appendix A). For the other cases, the ground state\nis not degenerate, but we apply the same weak field for\ncomparison.\nIII. RESULTS\nA. Phase diagram and static spin correlations\nBefore going into the spin dynamics, we discuss the\nground states of the clusters with edges shown in Fig. 1.\nFigures 2(a) and 2(b) show the αdependences of the\nsecond derivative of the ground-state energy for the sys-\ntems with the armchair and zigzag edges, respectively.\nThe two peaks at α=αc1andα=αc2indicate phase\ntransitions between the Kitaev QSL and the magnet-\nically ordered phases. Figures 2(c) and 2(d) display\nthe static interedge spin correlations defined by Eq. (7).\nFrom these data, we identify three different phases: the\nferromagnetic phase with a positive spin correlation for\nα < α c1, the Kitaev QSL with almost zero correlation\n(e) armchair (f) zigzag\niinioutiin\niout\nFIG. 2. αdependences of the second derivative of the ground-\nstate energy for the systems with (a) the armchair edges and\n(b) the zigzag edges. Corresponding static interedge spin cor-\nrelations are plotted in (c) and (d). (e) and (f) show the\nschematic pictures of the stripy order in the case of armchair\nand zigzag edges, respectively.\nforαc1< α < α c2, and the stripy phase with a nega-\ntive (positive) correlation for the system with the arm-\nchair (zigzag) edges for α > α c2. The antiferromagnetic\nand ferromagnetic spin correlations in the stripy phase\nare understood from the schematic pictures in Figs. 2(e)\nand 2(f), respectively. We note that in the ferromagnetic\nand stripy phases the spin correlations are dominant in a\nspecific spin component Szdue to the presence of edges,\nexcept for α=πwhere the ground state is degenerate in\nthe absence of the weak magnetic field hs.\nOur phase diagrams obtained for the clusters with\nedges are nearly identical to that for the same size clus-\nter under the periodic boundary conditions24. For the\nsystem with the armchair (zigzag) edges, we find that\nthe phase boundary between the ferromagnetic and Ki-\ntaev QSL phases is at αc1≃1.41π(1.49π) and that be-4\ntween the Kitaev QSL and stripy phases is at αc2≃1.57π\n(1.65π). These estimates are close to those for the clus-\nters with the periodic boundary conditions, αc1≃1.40π\nandαc2≃1.58π24. This indicates that the bulk proper-\nties are not much affected by the introduction of edges\neven for clusters of this size. In the following sections,\nwe will compute the spin dynamics in the three phases:\nFor the ferromagnetic, Kitaev QSL, and stripy phases,\nwe take α= 1.25π, 1.52π, and 1 .67π, respectively, for\nboth cases of the armchair and zigzag edges.\nLet us comment on the symmetry of the two clusters\nin Fig. 1. In the bulk system of the Kitaev-Heisenberg\nmodel, there is a four-sublattice transformation which\ndoes not change the form of the Hamiltonian with re-\nplacing KandJbyK+Jand−J, respectively24. This\ntransformation leads to the relation between the phase\nboundaries as tan αc2=−tanαc1−1. In the system\nwith the armchair edges, the relation holds for our es-\ntimates of αc1andαc2, since the cluster respects the\nfour-sublattice symmetry. In contrast, in the case of the\nzigzag edges, the symmetry is lost, and αc1andαc2do\nnot satisfy the relation.\nB. Real-time spin dynamics\nWe now turn to discuss how the spin at the edge is po-\nlarized when the AC magnetic field is applied to the spin\nat the other edge. Below, we present the results for the\nsystems with armchair and zigzag edges in Secs. III B 1\nand III B 2, respectively, and discuss the interedge spin\nresonance in the Kitaev QSL in Sec. III B 3.\n1. Armchair edge\nFigure 3 displays the time evolution of spin polariza-\ntion at the ioutth site in Eq. (6) for the system with\narmchair edges in Fig. 1(a). In the main panels, we show\nthe results for the period of the oscillating field, T= 10,\n20, 40, 60, and 80. Meanwhile, in the insets, we plot the\nFourier transformed spectra obtained by\nSν(ω) =\f\f\f\f2\ntmaxZtmax\n0Sν\niout(t)eiωtdt\f\f\f\f, (10)\nwhere we take tmax= 500 so that the lowest-energy scale\n2π/tmax∼0.013 is well below the excitation energy of\nthe localized fluxes ( ∼0.07) in the pure Kitaev model5.\nWe first discuss the results in the ferromagnetic phase\nshown in Figs. 3(a)–3(c). In this case, both Sx\niout(t) and\nSy\niout(t) show considerable oscillations, while Sz\niout(t) does\nnot. These behaviors are understood from the spin order-\ning in the ground state: As shown in Fig. 2(c), the spins\nare ferromagnetically ordered in the zdirection, for which\nfluctuations appear dominantly in the transverse compo-\nnents, Sx\nioutandSy\niout, rather than the longitudinal one\nSz\niout. In the Fourier transformed spectra shown in the in-\nsets, we find that the dominant Sx(ω) and Sy(ω) alwaysshow a peak around ω= Ω = 2 π/T. This result indicates\nthat the spin polarization is induced dominantly at the\nsame frequency of the input AC magnetic field.\nNext, we turn to the results in the Kitaev QSL phase\nshown in Figs. 3(d)–3(f). In contrast to the above ferro-\nmagnetic case, we find that only Sy\niout(t) shows consider-\nable oscillations, while the others do not. This behavior\ncan be understood from the fractional excitations in the\nKitaev QSL as follows. In the exact solution for the pure\nKitaev model, as mentioned in Sec. I, the spins are frac-\ntionalized into itinerant Majorana fermions and localized\nfluxes. The former has gapless excitations, while the lat-\nter is gapped5. The spin excitation is a composite of these\ntwo, and hence gapped. Indeed, the spin excitations by\nˆSxorˆSzat the iinorioutth site are gapped since these\nspin operators do not commute with the flux operators\ndefined by products of six spins on the hexagons includ-\ning the iinorioutth site5. This suppresses Sx\niout(t) and\nSz\niout(t) in Figs. 3(d) and 3(f), respectively. In contrast,\nˆSyat the iinorioutth site commutes with the flux op-\nerators, since the hexagons lack the ybond. In addition\nto the hexagonal fluxes, in the cluster with the armchair\nedges, there are additional flux operators defined only\nby the edge spins. While ˆSyat the iinorioutth site do\nnot commute with these fluxes, the spin excitations re-\nmain gapless because of the degeneracy in the ground\nstate (see Appendix A). These allow the excitation by\nˆSy\niinto yield long-range spin propagation via the gapless\nitinerant quasiparticles and induce Sy\niout(t) in Fig. 3(e).\nAlthough the above argument is valid only for the pure\nKitaev case, similar behavior is expected to appear in\nthe Kitaev QSL phase in the presence of weak Heisen-\nberg interactions. This is the reason why only Sy\niout(t)\nshows significantly large oscillations in Figs. 3(d)–3(f).\nThe resonant behaviors in the Kitaev QSL phase ex-\nhibit the following characteristics. First, while Sy(ω) al-\nways shows a peak around the input frequency Ω as in\nthe ferromagnetic case, the peak height does not decrease\nbut rather increases with ω, as shown in the inset of\nFig. 3(e). This characteristic behavior will be discussed\nin Sec. III B 3. Second, we note that a weak Heisenberg\ninteraction is crucial for the long-range spin propagation\nsince the ground state degeneracy in the pure Kitaev case\nwith α/π= 1.5 prohibits the propagation, as we will dis-\ncuss in detail in Sec. III C. Finally, the above argument\nalso allows static interedge spin correlation in the ydi-\nrection, Syy\nedge, also to develop, but it is almost zero as\nshown in Fig. 2(c). This indicates that the interedge spin\ncorrelations in the Kitaev QSL can only be dynamically\nenhanced.\nLastly, we discuss the results in the stripy phase shown\nin Figs. 3(g)–3(i). In this case, the results are similar to\nthe ferromagnetic case in Figs. 3(a)–3(c). The reason\nis common: As shown in Fig. 2(c), the spins are anti-\nferromagnetically ordered in the zdirection in this stripy\nphase, and hence, the transverse components Sx\niout(t) and\nSy\niout(t) are induced dominantly at the input frequency.5\n−0.2−0.10.00.10.20.30.4(a)α/π =1.25 (ferro)\nSx\niout(t)\nT=10\nT=20\nT=40\nT=60\nT=80(b)α/π =1.25 (ferro)\nSy\niout(t)(c)α/π =1.25 (ferro)\nSz\niout(t)\n−0.2−0.10.00.10.20.30.4(d)α/π =1.52 (QSL)\nSx\niout(t)\nT=10\nT=20\nT=40\nT=60\nT=80(e)α/π =1.52 (QSL)\nSy\niout(t)(f)α/π =1.52 (QSL)\nSz\niout(t)\n0 100 200 300 400 500\nt−0.2−0.10.00.10.20.30.4(g)α/π =1.67 (stripy)\nSx\niout(t)\nT=10\nT=20\nT=40\nT=60\nT=80\n0 100 200 300 400 500\nt(h)α/π =1.67 (stripy)\nSy\niout(t)\n0 100 200 300 400 500\nt(i)α/π =1.67 (stripy)\nSz\niout(t)0.00 0.25 0.50 0.75\nω0.000.050.10\nSx(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSy(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSz(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSx(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSy(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSz(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSx(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSy(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSz(ω)\nFIG. 3. Time evolution of the spin polarization Sν\niout(t) in Eq. (6) for the system with armchair edges in (a)–(c) the ferromagnetic\nphase at α/π = 1.25, (d)–(f) the Kitaev QSL phase at α/π = 1.52, and (g)–(i) the stripy phase at α/π = 1.67: (a),(d),(g)\nν=x, (b),(e),(h) ν=y, and (c),(f),(i) ν=z. The data for the input oscillation periods T= 10, 20, 40, 60, and 80 are shown.\nThe insets display the corresponding Fourier transformed spin polarizations in Eq. (10). The vertical dashed lines denote the\nfrequencies corresponding to the values of T,ω= Ω = 2 π/T.\n2. Zigzag edge\nFigure 4 displays the results for the system with the\nzigzag edges in Fig. 1(b), obtained by the same condi-\ntions for the armchair case. For the ferromagnetic andstripy phases shown in Figs. 4(a)–4(c) and 4(g)–4(i), re-\nspectively, we find similar tendency to the armchair case:\nThe spin polarizations in the xandydirections are in-\nduced significantly, while that in the zdirection is rather\nsuppressed. This is again understood from the fact that6\n−0.2−0.10.00.10.20.30.4(a)α/π =1.25 (ferro)\nSx\niout(t)\nT=10\nT=20\nT=40\nT=60\nT=80(b)α/π =1.25 (ferro)\nSy\niout(t)(c)α/π =1.25 (ferro)\nSz\niout(t)\n−0.2−0.10.00.10.20.30.4(d)α/π =1.52 (QSL)\nSx\niout(t)\nT=10\nT=20\nT=40\nT=60\nT=80(e)α/π =1.52 (QSL)\nSy\niout(t)(f)α/π =1.52 (QSL)\nSz\niout(t)\n0 100 200 300 400 500\nt−0.2−0.10.00.10.20.30.4(g)α/π =1.67 (stripy)\nSx\niout(t)\nT=10\nT=20\nT=40\nT=60\nT=80\n0 100 200 300 400 500\nt(h)α/π =1.67 (stripy)\nSy\niout(t)\n0 100 200 300 400 500\nt(i)α/π =1.67 (stripy)\nSz\niout(t)0.00 0.25 0.50 0.75\nω0.000.050.10\nSx(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSy(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSz(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSx(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSy(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSz(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSx(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSy(ω)\n0.00 0.25 0.50 0.75\nω0.000.050.10\nSz(ω)\nFIG. 4. Corresponding plots to Fig. 3 for the system with zigzag edges.\nthe ordered moments in each phase appear in the zdi-\nrection as shown in Fig. 2(d).\nMeanwhile, in the Kitaev QSL phase, as shown in\nFigs. 4(d)–4(f), we find that Sz\niout(t) shows significant\nlarge oscillations, while Sx\niout(t) and Sy\niout(t) do not. This\ncan also be understood from the flux excitations dis-\ncussed above for the armchair case. In the current case,\nthe zigzag edges lack the zbonds, and hence, the Szcom-\nponents can be excited without the gapped excitationsowing to the ground state degeneracy; see Appendix A.\nInterestingly, the amplitude of Sz\niout(t) varies nonmono-\ntonically with Tand takes the maximum value around\nT= 40, as shown in Fig. 4(f). In this case also, we ob-\nserve the peaks in Sz(ω) at almost the input frequencies,\nas shown in the inset of Fig. 4(f), while the peak heights\nshow nonmonotonic ωdependence. These features will\nbe discussed in the next section.7\n3. Interedge resonance\nIn Fig. 5, we show Ω = 2 π/Tdependences of the max-\nimum values of Sν(ω) (ν=x, y, z ) within the range\nof 2/tmax < ω/ 2π≤0.5, which are represented as\nSν(ωpeak), in the systems with (a)-(c) the armchair edges\nand (d)-(f) the zigzag edges. We also plot the values of\nωpeakas functions of Ω in each inset. Here, we choose\nthe lower limit as twice 2 π/tmaxto avoid an artifact near\n2π/tmaxand the upper limit to be sufficiently larger than\nthe bandwidth of the dynamical spin correlation26.\nIn both armchair and zigzag cases, Sx(ωpeak) and\nSy(ωpeak) have large values in the ferromagnetic and\nstripy phases, while Sz(ωpeak) are suppressed, since these\nphases show the spin orderings along the zdirection as\nmentioned above. The values of Sx(ωpeak) and Sy(ωpeak)\ndecrease as Ω increases. In addition, we find that the val-\nues of ωpeakare close to Ω as shown in each inset. These\nare indications of conventional magnetic resonances; the\nspin polarization is induced dominantly at the input fre-\nquency, as long as the magnetic excitations are available\nin the frequency range.\nIn contrast, the interedge spin resonance behaves dif-\nferently in the Kitaev QSL phase. First of all, we find\nthat the dominant polarizations, Sy(ωpeak) for the arm-\nchair case and Sz(ωpeak) for the zigzag case, show broad\npeaks in the wide frequency range up to Ω ≃1.5, while\nthe latter shows a sharp peak at a smaller Ω ≃0.2 as well.\nFor both cases, the relation ωpeak∼Ω holds, except for\nsmall Ω (Ω ≲0.2 for the armchair case and Ω ≲0.1 for\nthe zigzag case); the deviation might be due to the finite-\nsize effect. The broad responses with ωpeak∼Ω, as well\nas the sharp peak in the zigzag case, can be ascribed to\nthe itinerant Majorana fermions, whose density of states\nshows a continuum up to ω= 1.55. This point will be\nfurther discussed in Sec. III C.\nIn addition, for the other spin components with sup-\npressed polarizations, we find that ωpeakis almost con-\nstant irrespective of Ω27. This behavior could be ex-\nplained by the flux gap that governs the low-energy ex-\ncitations in these spin components. We note that the\nconstant values of ωpeak∼0.2 in the armchair case is\nlarger than the energy of the low-energy coherent peak\natω∼0.1 in the dynamical spin structure factor of the\npure Kitaev model in the thermodynamic limit26, but\nthis might also be due to the finite-size effect.\nIn the Kitaev QSL, Sx(ωpeak) is larger than Sz(ωpeak)\nin the armchair case, while Sx(ωpeak) and Sy(ωpeak) are\nalmost the same in the zigzag case. This is understood\nfrom the geometry of the Kitaev bonds in each cluster. In\nthe armchair case, as shown in Fig. 1(a), the zbonds on\nwhich ˆSzcomponents interact via the Kitaev interaction\nare along the edges and perpendicular to the direction\nfrom iintoiout, which may suppress the interedge spin\ntransport of the zcomponent. Meanwhile, in the zigzag\ncase shown in Fig. 1(b), both xandybonds are along the\nedges and related with each other by symmetry, leading\nto almost the same interedge resonances assisted by thezbonds connecting them.\nC. Comparison with dynamical spin correlations\nLet us discuss the characteristic interedge spin reso-\nnances in comparison with the dynamical interedge spin\ncorrelations Cνν\nedge(ω) defined in Eq. (8). Figure 6 displays\nCνν\nedge(ω) for the armchair and zigzag cases while varying\nα. In the following, we show that Cνν\nedge(ω) explains well\nthe intensities and Ω dependences of the induced spin\npolarizations in Fig. 5.\nIn the system with armchair edges, Cxx\nedge(ω) and\nCyy\nedge(ω) show considerable intensities over the broad ω\nrange, whereas Czz\nedge(ω) is almost zero except for the\nlow-ωweights near the phase boundaries at α=αc1and\nαc2. In the ferromagnetic phase for α≤αc1and the\nstripy phase for α≥αc2, this is again consistent with the\nfact that the spin moments are ordered along the zdirec-\ntion. A striking difference between Cxx\nedge(ω) and Cyy\nedge(ω)\nappears in the Kitaev QSL phase for αc1≤α≤αc2;\nCyy\nedge(ω) has large spectral weights over the broad ω\nrange, while Cxx\nedge(ω) is almost zero. Notably, the inten-\nsity of Cyy\nedge(ω) is stronger than those in the ferromag-\nnetic and stripy phases, while it vanishes for the pure\nKitaev case at α/π= 1.5 because of the degeneracy in\nthe ground state (see Appendix B). This strong Cyy\nedge(ω)\nexplains well the broad response in Sy(ωpeak) found in\nFig. 5(b). In addition, we note that Cxx\nedge(ω) has weak\nintensities at low ω∼0.1, as shown in Fig. 6(a). This\nalso explains well the small peak in Sx(ωpeak) found in\nFig. 5(a).\nMeanwhile, in the system with zigzag edges, Cνν\nedge(ω)\nin the ferromagnetic and stripy phases behave qualita-\ntively similarly to those in the armchair case. In the\nKitaev QSL phase, however, strong intensity appears in\nCzz\nedge(ω) over the broad ωrange, while Cxx\nedge(ω) and\nCyy\nedge(ω) are almost absent. Again, this explains well the\nbroad response in Sz(ωpeak) found in Fig. 5(f). Further-\nmore, the sharp peak at ω∼0.2 in Fig. 5(f) is also con-\nsistent with the strong intensity of Czz\nedge(ω) in Fig. 6(f).\nThe interedge resonances in the broad ωrange in the\nKitaev QSL phase are mediated by the itinerant Majo-\nrana fermions whose excitation spectrum has a contin-\nuum in the broad energy range. This is explicitly shown\nby calculating the dynamical spin correlations for the\npure Kitaev model at α/π = 1.5 by using the Majo-\nrana representation, which we denote CMaj\nedge(ω); see Ap-\npendix B for the details of the calculations. Figure 7\nshows the results of CMaj\nedge(ω) in comparison with Cyy\nedge(ω)\nandCzz\nedge(ω) around α/π= 1.5. Note that here we com-\npare their absolute values since the sign of CMaj\nedge(ω) in\nEq. (B3) is not well defined. We find that the broad\nresponses of Cyy\nedge(ω) and Czz\nedge(ω) in the vicinity of\nα/π= 1.5 appear in the same energy range of CMaj\nedge(ω)\nwith showing similar ωdependences. This indicates that8\n0.000.020.040.060.080.10\narmchair, Sx(ωpeak) (a)\nα=1.25 (ferro)\nα=1.52 (QSL)\nα=1.67 (stripy)\narmchair, Sy(ωpeak) (b)\n armchair, Sz(ωpeak) (c)\n0.0 0.5 1.0 1.5 2.0\nΩ0.000.020.040.060.080.10\nzigzag, Sx(ωpeak) (d)\nα=1.25 (ferro)\nα=1.52 (QSL)\nα=1.67 (stripy)\n0.0 0.5 1.0 1.5 2.0\nΩ\nzigzag, Sy(ωpeak) (e)\n0.0 0.5 1.0 1.5 2.0\nΩ\nzigzag, Sz(ωpeak) (f)0.0 0.5 1.0 1.5\nΩ0.00.51.01.5ωpeak\n0.0 0.5 1.0 1.5\nΩ0.00.51.01.5ωpeak\n0.0 0.5 1.0 1.5\nΩ0.00.51.01.5ωpeak\n0.0 0.5 1.0 1.5\nΩ0.00.51.01.5ωpeak\n0.0 0.5 1.0 1.5\nΩ0.00.51.01.5ωpeak\n0.0 0.5 1.0 1.5\nΩ0.00.51.01.5ωpeak\nFIG. 5. Maximum values of the Fourier transformed spin polarizations Sν(ωpeak) for the input oscillating magnetic field with\nΩ = 2 π/T in the systems with (a)–(c) the armchair edges and (d)–(f) the zigzag edges: (a) and (d) ν=x, (b) and (e) ν=y,\nand (c) and (f) ν=z. The insets display Ω dependences of ωpeak. The gray dashed line shows the relation ωpeak= Ω.\nthe broad responses in the Kitaev QSL phase are domi-\nnated by the itinerant Majorana excitations.\nWhile our results are limited to the small clusters, we\nexpect that the interedge resonances appear also in larger\nsystems since the itinerant Majorana fermions propagate\nover long distances in the Kitaev QSL. This is demon-\nstrated by calculating |CMaj\nedge(ω)|while changing the sys-\ntem width. Figure 8 shows the maximum intensity of\n|CMaj\nedge(ω)|as a function of the number of the unit cells\nin the direction perpendicular to the edges, L⊥. We find\nthat the dynamical correlations decay slowly: the zigzag\ncase roughly obeys ∝1/L⊥, while the armchair case\nshows crossover from ∝1/L⊥to∝1/L3\n⊥. The results\nappear to be consistent with the Majorana-mediated spin\ncorrelations17,22. Thus, we believe that, when domi-\nnated by the itinerant Majorana fermions, the interedge\ndynamical spin correlations become long-range in realspace, even in the presence of weak Heisenberg interac-\ntions28–30.\nCombining these results with the almost constant be-\nhaviors of ωpeakirrespective of Ω for the other suppressed\ncomponents in Fig. 5, we conclude that the interedge\nspin resonances in the Kitaev QSL are good probes of\ntwo types of fractional excitations, itinerant Majorana\nfermions and localized fluxes. The resonance in the spin\ncomponent which does not excite the fluxes on hexagons\nleads to broad responses with ωpeak∼Ω, as found in\nFigs. 5(b) and 5(f). This is a clear indication of the itin-\nerant Majorana excitations. Meanwhile, the responses in\nthe other spin components appear around a small con-\nstant ωpeak. This is an indication of the gapped flux\nexcitations. We emphasize that weak Heisenberg interac-\ntions are essential for the interedge spin resonances since\nallCνν\nedge(ω) vanish for the pure Kitaev model because of9\n012ωCxx\nedge(ω)armchairαc1αc2\n(a)\nCxx\nedge(ω)zigzagαc1αc2\n(b)\n012ωCyy\nedge(ω)(c)\nCyy\nedge(ω)(d)\n1.00 1.25 1.50\nα/π012ωCzz\nedge(ω)(e)\n1.00 1.25 1.50\nα/πCzz\nedge(ω)(f)\n−0.50.00.5\n−0.50.00.5\n−0.50.00.5\nFIG. 6. Dynamical interedge spin correlations Cνν\nedge(ω)\n[Eq. (8)] for (a),(c),(e) the armchair edge and (b),(d),(f) the\nzigzag edge: (a) and (b) ν=x, (c) and (d) ν=y, and (e)\nand (f) ν=z.\nthe ground-state degeneracy (see Appendix B).\nIV. SUMMARY\nIn summary, we have studied how an AC local mag-\nnetic field at an edge of the system induces spin polar-\nizations at the opposite edge in the Kitaev-Heisenberg\nmodel with ferromagnetic Kitaev interactions by using\nthe exact diagonalization. We found that in the Kitaev\nQSL phase the spin polarizations are resonantly induced\nin a particular spin component, in stark contrast to the\nmagnetically ordered phases where conventional mag-\nnetic resonances appear in the transverse spin compo-\nnents. The spin resonance in the Kitaev QSL shows the\nfollowing peculiar features, stemming from the fraction-\nalization of spin degree of freedom into two types of frac-\ntional excitations, itinerant Majorana fermions and local-\nized fluxes: (i) It appears dominantly in the spin compo-\nnent which does not excite flux excitations, (ii) the dom-\ninant resonance appears in a broad range of frequency,\nreflecting the continuum of Majorana excitations, (iii) it\nis accompanied by subdominant resonances in the other\nspin components at a small constant frequency corre-\n0.00.20.40.60.81.01.21.41.6ω(a)armchair\n1.451.501.55\nα/π(b)\n (c)zigzag\n1.451.501.55\nα/π(d)αc1\n0.00.10.20.30.40.5FIG. 7. Comparison between (a) |CMaj\nedge(ω)|calculated by\nEq. (B3) for the pure Kitaev model at α/π = 1.5 and (b)\nan enlarged plot of |Cyy\nedge(ω)|around α/π= 1.5 for the sys-\ntem with the armchair edges. (c) and (d) The corresponding\nplots for the zigzag case, where |Czz\nedge(ω)|is plotted in (d).\n100101102\nL⊥10−210−1100|CMaj\nedge(ω)|max\narmchair L/bardbl=4\nzigzag L/bardbl=3\nFIG. 8. Maximum of |CMaj\nedge(ω)|as a function of the number\nof the unit cells in the direction perpendicular to the edges,\nL⊥. The data are calculated for the clusters with the numbers\nof the unit cell along the edge, L∥= 4 and 3, for the armchair\nand zigzag cases, respectively; the result for the smallest L⊥\nin each case corresponds to that in Figs. 7(a) and 7(c).\nsponding to the flux excitation gap, (iv) both resonances\nvanish in the exact Kitaev QSL because of the ground-\nstate degeneracy and require weak Heisenberg interac-\ntions, and (v) they are induced only dynamically, despite\nthe disappearance of the static spin correlations. These\nresults elucidate that the nonlocal spin dynamics in the\nwide frequency range contains information on both two10\ntypes of fractional excitations in the Kitaev QSL, which\ncannot be captured by the spin transport in the low-\nenergy limit in the previous studies22,23. While our cal-\nculations were done for small size clusters, the interedge\nresonance is expected to be observed in larger systems,\nsince it is mediated by itinerant Majorana excitations\nthat propagate over long distances. These results indi-\ncate that the interedge dynamical spin resonance is useful\nfor probing the two types of fractional excitations in the\nKitaev QSL, which are usually difficult to observe only\nfrom static physical quantities.\nA straightforward experiment would be implemented\nby using a scanning tunneling microscope (STM) tip with\nmagnetic atoms or the atomic force microscopy (AFM)\nto apply an AC magnetic field at the edge and measure\nthe spin polarization at the opposite edge. This could be\nperformed, for example, for a thin flake of a candidate\nmaterial α-RuCl 3. Similar experiments would be pos-\nsible in interface or heterostructure of a Kitaev magnet\nand a ferromagnetic material, where the AC magnetic\nfield can be applied to the edge spins by the ferromag-\nnetic resonance. Careful analysis of the dynamics in each\nspin component and its dependence on the edge structure\nwould pave the way for creating and controlling the frac-\ntional excitations in the Kitaev QSL through the spin\ndegree of freedom.\nACKNOWLEDGMENTS\nWe wish to thank Y. Kato, K. Fukui, and T. Okubo\nfor fruitful discussions. This work was also supported\nby the National Natural Science Foundation of China\n(Grant No. 12150610462). TM was supported by Build-\ning of Consortia for the Development of Human Re-\nsources in Science and Technology, MEXT, Japan. This\nwork was supported by Grant-in-Aid for Scientific Re-\nsearch Nos. 19H05825, JP19K03742, and 20H00122 from\nthe Ministry of Education, Culture, Sports, Science\nand Technology, Japan. It is also supported by JST\nCREST Grant No. JPMJCR18T2 and JST PRESTO\nGrant No. JPMJPR19L5.\nAppendix A: Degeneracy in the pure Kitaev model\nwith edges\nIn this Appendix, we show that the ground state of the\npure Kitaev model with α= 3π/2 has the degeneracy\nin both cases of the armchair and the zigzag edges. In\nthe pure Kitaev model, one can define the flux operator\nˆWpby a product of six spins for each hexagon p, which\ncommutes with the Hamiltonian5. We show the examples\nin Fig. 9; note that ˆWpin (a) [(b)] commutes with ˆSy\niin\n(ˆSz\niin), since the hexagon lacks the y(z) bond at the iinth\nsite, as discussed in Sec. III B 1 (III B 2). In addition to\nthe six-spin flux operators, at the edges of the system\nthere are additional flux operators defined only by the\n(a) armchair (b) zigzag\nxy\nz\n01 2\n345 6\n7402\n513\nWp\nWout4WpWout3a\nWout3b\n^^\n^^\n^uij = −1\nuiiniout\nWout2^uiiniouty\nzzFIG. 9. Schematic pictures of the flux operators for (a) the\narmchair edges and (b) the zigzag edges. The numbers denote\nthe sites used for the definitions of the fluxes in Eq. (A1) and\nEq. (A2). We show examples of the six-spin flux operator\nˆWp, four-spin ( ˆW4) and two-spin flux operators ( ˆW2) and\nthe three-spin flux operators ( ˆW3aand ˆW3b). We also show\nthe interedge correlations of the localized Majorana particles\n(uiiniout) by the purple lines. In (a), we represent the zbonds\nwhere uijtakes -1 with red dashed lines. See Appendix B for\nuiinioutanduij.\nedge spins. For instance, the flux operators including the\noutput site are given by\nˆW4\nout= 24ˆSz\n0ˆSy\n1ˆSx\n2ˆSz\n3,ˆW2\nout= 22ˆSx\n0ˆSy\n7, (A1)\nfor the armchair case, and\nˆW3a\nout= 23ˆSx\n0ˆSz\n1ˆSy\n2,ˆW3b\nout= 23ˆSy\n0ˆSz\n5ˆSx\n4, (A2)\nfor the zigzag case; see Fig. 9. Since these flux operators\nˆWq(q= 4, 2, 3 a, and 3 b) commute with the Hamiltonian\natα= 3π/2, the ground state |ΦGS⟩is the eigenstate of\nthe flux operators as\nˆWq\nout|ΦGS⟩=W|ΦGS⟩, (A3)\nwhere the eigenvalue Wtakes +1 or −1. Meanwhile, all\nthe eigenstates of the Hamiltonian can be taken to be\nreal since the Kitaev Hamiltonian does not include the\ncomplex elements in the conventional basis set composed\nof the eigenstates of ˆSz\ni. Therefore, if the ground state\n|ΦGS⟩is unique, we obtain ⟨ΦGS|ˆWq\nout|ΦGS⟩= 0 since\nˆWq\noutis the pure imaginary operator including a single\nˆSy\ni. This contradicts with Eq. (A3), meaning that the\nassumption of a unique ground state is incorrect. Hence,\nthe ground state of the pure Kitaev model with the arm-\nchair and the zigzag edges must be degenerate. For the\n24-site clusters shown in Fig. 1, we numerically confirm\nthat the ground state has eightfold (fourfold) degener-\nacy for the clusters with the armchair (zigzag) edges.\nWe note that the numbers of the degenerate states can\nbe accounted for by the numbers of independent flux-\ntype operators traversing the system from one edge to\nthe other and those consisting of edge spins.11\nAppendix B: Dynamical spin correlations in the\npure Kitaev model\nIn this Appendix, we describe the method to calculate\nthe dynamical spin correlations for the pure Kitaev model\nin Figs. 7(a) and 7(c). We adopt the Majorana represen-\ntation of the Hamiltonian in Eq. (1) at α= 3π/2, which\nis given by5\nˆHK=1\n4X\n⟨i,j⟩νuν\nijiˆciˆcj, (B1)\nwhere the spin operator is represented as ˆSν\ni=i\n2ˆbν\niˆciby\nintroducing four Majorana fermions {ˆci,ˆbx\ni,ˆby\ni,ˆbz\ni}. Here,\nuν\nij=⟨ΦGS|ˆuν\nij|ΦGS⟩=⟨ΦGS|iˆbν\niˆbν\nj|ΦGS⟩; ˆuν\nijcommutes\nwith the Hamiltonian and uν\nijtakes±1.\nIn this Majorana representation, the interedge dynam-ical spin correlation is given by\n⟨ΦGS|ˆSν\niin(t)ˆSν\niout|ΦGS⟩\n=−1\n4uν\niiniout⟨ΦGS|iˆciin(t)ˆciout|ΦGS⟩, (B2)\nwhere ν=yandzfor the armchair and zigzag case, re-\nspectively, and uν\niinioutis defined for the unpaired ˆbν\niinand\nˆbν\niouton the edges as uν\niiniout=⟨ΦGS|iˆbν\niinˆbν\niout|ΦGS⟩. The\ncorrelations for the other spin components vanish. By\nsubstituting Eq. (B2) to Eq. (8), we obtain the dynam-\nical spin correlation between the edges in the Majorana\nrepresentation as\nCMaj\nedge(ω) =−uν\niiniout\n8πZ∞\n−∞⟨ΦGS|iˆciin(t)ˆciout|ΦGS⟩eiωtdt,\n(B3)\nwhere the time-dependent operator is defined by ˆHKin\nwhich uν\nijare chosen to realize the flux-free ground state\n|Φ0⟩: We take all uν\nij= +1 for the zigzag case, while\nwe flip uν\nijto−1 on the zbonds in one column for the\narmchair case as shown in Fig. 9(a). In both cases, how-\never, the sign of CMaj\nedge(ω) is indefinite due to the fac-\ntor of uν\niiniout; we plot the absolute value |CMaj\nedge(ω)|in\nFigs. 7 and 8, which corresponds to setting uν\niiniout=−1\nin Eq. (B3). Note that Eq. (B3) besides this factor cor-\nresponds to the propagator of the Majorana fermions ˆ ci.\n1E. Majorana, “Teoria simmetrica dell’elettrone e del\npositrone,” Il Nuovo Cimento 14, 171 (1937).\n2F. Wilczek, “Majorana returns,” Nature Phys 5, 614\n(2009).\n3A. Kitaev, “Fault-tolerant quantum computation by\nanyons,” Annals of Physics 303, 2–30 (2003).\n4M. Freedman, A. Kitaev, M. Larsen, and Z. Wang, “Topo-\nlogical quantum computation,” Bulletin of the American\nMathematical Society 40, 31 (2003).\n5A. Kitaev, “Anyons in an exactly solved model and be-\nyond,” Annals of Physics 321, 2 (2006).\n6G. Jackeli and G. Khaliullin, “Mott Insulators in the\nStrong Spin-Orbit Coupling Limit: From Heisenberg to a\nQuantum Compass and Kitaev Models,” Phys. Rev. Lett.\n102, 017205 (2009).\n7J. Chaloupka, G. Jackeli, and G. Khaliullin, “Kitaev-\nHeisenberg Model on a Honeycomb Lattice: Possible Ex-\notic Phases in Iridium Oxides A2IrO3,” Phys. Rev. Lett.\n105, 027204 (2010).\n8R. Yadav, N. A. Bogdanov, Vamshi M. Katukuri, S. Nishi-\nmoto, J. Van Den Brink, and L. Hozoi, “Kitaev exchange\nand field-induced quantum spin-liquid states in honeycomb\nα-RuCl 3,” Scientific reports 6, 1 (2016).\n9Y. Motome and J. Nasu, “Hunting Majorana Fermions in\nKitaev Magnets,” Journal of the Physical Society of Japan\n89, 012002 (2020).\n10Y. Kasahara, T. Ohnishi, Y. Mizukami, O. Tanaka, S. Ma,\nK. Sugii, N. Kurita, H. Tanaka, J. Nasu, Y. Motome,T. Shibauchi, and Y. Matsuda, “Majorana quantization\nand half-integer thermal quantum Hall effect in a Kitaev\nspin liquid,” Nature 559, 227 (2018).\n11M. Yamashita, J. Gouchi, Y. Uwatoko, N. Kurita, and\nH. Tanaka, “Sample dependence of half-integer quantized\nthermal Hall effect in the Kitaev spin-liquid candidate α-\nRuCl 3,” Phys. Rev. B 102, 220404(R) (2020).\n12T. Yokoi, S. Ma, Y. Kasahara, S. Kasahara, T. Shibauchi,\nN. Kurita, H. Tanaka, J. Nasu, Y. Motome, C. Hickey,\nS. Trebst, and Y. Matsuda, “Half-integer quantized\nanomalous thermal Hall effect in the Kitaev material can-\ndidate α-RuCl 3,” Science 373, 568 (2021).\n13J. A. N. Bruin, R. R. Claus, Y. Matsumoto, N. Kurita,\nH. Tanaka, and H. Takagi, “Robustness of the thermal\nHall effect close to half-quantization in α-RuCl 3,” Nature\nPhysics 18, 401 (2022).\n14P. Czajka, T. Gao, M. Hirschberger, P. Lampen-Kelley,\nA. Banerjee, N. Quirk, D. G. Mandrus, S. E. Nagler,\nand N. P. Ong, “Planar thermal Hall effect of topological\nbosons in the Kitaev magnet α-RuCl 3,” Nature Materials\n, 1 (2022).\n15´E. Lefran¸ cois, G. Grissonnanche, J. Baglo, P. Lampen-\nKelley, J.-Q. Yan, C. Balz, D. Mandrus, S. E. Nagler,\nS. Kim, Y.-J. Kim, N. Doiron-Leyraud, and L. Taillefer,\n“Evidence of a Phonon Hall Effect in the Kitaev Spin Liq-\nuid Candidate α-RuCl 3,” Phys. Rev. X 12, 021025 (2022).\n16A. J. Willans, J. T. Chalker, and R. Moessner, “Site di-\nlution in the Kitaev honeycomb model,” Phys. Rev. B 84,12\n115146 (2011).\n17A. Koga, Y. Murakami, and J. Nasu, “Majorana corre-\nlations in the Kitaev model with ordered-flux structures,”\nPhys. Rev. B 103, 214421 (2021).\n18G. Baskaran, S. Mandal, and R. Shankar, “Exact Re-\nsults for Spin Dynamics and Fractionalization in the Ki-\ntaev Model,” Phys. Rev. Lett. 98, 247201 (2007).\n19A. J. Willans, J. T. Chalker, and R. Moessner, “Disorder\nin a Quantum Spin Liquid: Flux Binding and Local Mo-\nment Formation,” Phys. Rev. Lett. 104, 237203 (2010).\n20D. Takikawa, M. G. Yamada, and S. Fujimoto, “Dissi-\npationless spin current generation in a Kitaev chiral spin\nliquid,” Phys. Rev. B 105, 115137 (2022).\n21M. O. Takahashi, M. G. Yamada, M. Udagawa,\nT. Mizushima, and S. Fujimoto, “Non-local spin corre-\nlation as a signature of Ising anyons trapped in vacancies\nof the Kitaev spin liquid,” arXiv preprint arXiv:2211.13884\n(2022).\n22T. Minakawa, Y. Murakami, A. Koga, and J. Nasu,\n“Majorana-Mediated Spin Transport in Kitaev Quantum\nSpin Liquids,” Phys. Rev. Lett. 125, 047204 (2020).\n23H. Taguchi, A. Koga, Y. Murakami, J. Nasu, and\nH. Tsuchiura, “Field-direction Dependence of Majorana-\nmediated Spin Transport,” JPS Conf. Proc. 38, 011152\n(2023).\n24J. Chaloupka, G. Jackeli, and G. Khaliullin, “Zigzag Mag-netic Order in the Iridium Oxide Na 2IrO3,” Phys. Rev.\nLett. 110, 097204 (2013).\n25M. Kawamura, K. Yoshimi, T. Misawa, Y. Yamaji,\nS. Todo, and N. Kawashima, “Quantum lattice model\nsolver HΦ,” Computer Physics Communications 217, 180\n(2017).\n26J. Knolle, D. L. Kovrizhin, J. T. Chalker, and R. Moess-\nner, “Dynamics of a Two-Dimensional Quantum Spin\nLiquid: Signatures of Emergent Majorana Fermions and\nFluxes,” Phys. Rev. Lett. 112, 207203 (2014).\n27The data for Sz(ωpeak) at Ω = 2 π/10≃0.63 and Sz(ωpeak)\nat Ω = 2 π/5≃1.26 in the armchair case deviate from the\nconstant behavior, and rather close to Ω. In these cases,\nwe also observe peaks around the constant values, but the\npeak heights are slightly smaller than those around Ω.\n28K. S. Tikhonov, M. V. Feigel’man, and A. Yu. Kitaev,\n“Power-law spin correlations in a perturbed spin model on\na honeycomb lattice,” Phys. Rev. Lett. 106, 067203 (2011).\n29S. Mandal, Subhro Bhattacharjee, K. Sengupta,\nR. Shankar, and G. Baskaran, “Confinement-\ndeconfinement transition and spin correlations in a\ngeneralized Kitaev model,” Phys. Rev. B 84, 155121\n(2011).\n30X.-Y. Song, Y.-Z. You, and L. Balents, “Low-Energy Spin\nDynamics of the Honeycomb Spin Liquid Beyond the Ki-\ntaev Limit,” Phys. Rev. Lett. 117, 037209 (2016)." }, { "title": "1204.0239v2.Spin_dynamics_in_the_Kapitza_Dirac_effect.pdf", "content": "Spin dynamics in the Kapitza-Dirac effect\nSven Ahrens,1Heiko Bauke,1,\u0003Christoph H. Keitel,1and Carsten Müller1, 2,y\n1Max-Planck-Institut für Kernphysik, Saupfercheckweg 1, 69117 Heidelberg, Germany\n2Institut für Theoretische Physik I, Heinrich-Heine-Universität Düsseldorf, Universitätsstraße 1, 40225 Düsseldorf, Germany\n(Dated: November 15, 2018)\nElectron spin dynamics in Kapitza-Dirac scattering from a standing laser wave of high frequency and high\nintensity is studied. We develop a fully relativistic quantum theory of the electron motion based on the time-\ndependent Dirac equation. Distinct spin dynamics, with Rabi oscillations and complete spin-flip transitions, is\ndemonstrated for Kapitza-Dirac scattering involving three photons in a parameter regime accessible to future\nhigh-power X-ray laser sources. The Rabi frequency and, thus, the di \u000braction pattern is shown to depend crucially\non the spin degree of freedom.\nPACS numbers: 03.65.Pm, 42.50.Ct, 42.55.Vc, 03.75.-b\nIntroduction The di \u000braction of an electron beam from\na standing wave of light is referred to as the Kapitza-Dirac\ne\u000bect [1,2]. This process points out the quantum wave nature\nof the electron and may be considered as an analogue of the\noptical di \u000braction of light on a grating, but with the roles of\nlight and matter interchanged. Predicted already in 1933, a\nclear experimental confirmation of the Kapitza-Dirac e \u000bect\nas originally proposed has been achieved only recently [ 3].\nIt has stimulated renewed theoretical interest in the process\n[4], advancing earlier treatments [ 5,6]. A related success-\nful experiment observed the (classical) scattering of electrons\nin a standing laser wave [ 7]. Both experiments operated at\nmoderate light intensities between 109W/cm2and1014W/cm2\nand near optical wavelengths. The existing theoretical stud-\nies, accordingly, have treated the electron quantum dynamics\nnonrelativistically. The nonrelativistic Kapitza-Dirac e \u000bect has\nbeen observed experimentally also using atomic beams [8].\nThe current development of novel light sources, envisaged\nto provide field intensities in excess of 1020W/cm2and field\nfrequencies in the hard X-ray domain [ 9,10], raises the ques-\ntion as to how the Kapitza-Dirac e \u000bect is modified in this\nhitherto unexplored parameter regime. This calls for a fully\nrelativistic treatment of the process within the Dirac theory,\nvalid at high electron energies and, in particular, accounting\nfor the electron spin. Relativistic considerations of quantum\nmechanical electron scattering in two counterpropagating light\nwaves were provided, but assuming nonequal laser frequen-\ncies and disregarding the electron spin degree of freedom [ 11].\nSpin signatures in the Kapitza-Dirac e \u000bect have so far been\nexamined within the nonrelativistic framework of the Pauli\nequation [ 12] and by solving the classical equations of motion\n[13]. Both studies found negligibly small spin e \u000bects. We note\nthat the influence of the electron spin has also been investi-\ngated with respect to free-electron motion [ 14], bound-electron\ndynamics [ 15], atomic photoionization [ 16], and Compton and\nMott [17] scattering in strong plane-wave laser fields.\nIn this Letter, we present a fully relativistic consideration\nof the Kapitza-Dirac e \u000bect within the framework of Dirac the-\nory. We focus on the relevance of spin-flip transitions when\nthe process occurs in X-ray laser fields of high intensity. Wedemonstrate that under suitable conditions, the di \u000braction prob-\nability depends considerably on the electron spin. Thus, the\nspin degree of freedom significantly influences the quantum\nmechanical scattering dynamics.\nRelativistic theory We consider a quantum electron\nwave packet in a standing linearly polarized light wave (see\nFig. 1) with maximal electric field strength E, wave vector k,\nand wavelength \u0015=2\u0019=jkj=2\u0019=k, respectively. The laser is\nmodeled by the vector potential\nA(x;t)=\u0000E\nkcos(k\u0001x) sin( ckt)w(t) (1)\nintroducing the speed of light cand the temporal envelope\nfunction w(t). The relativistic quantum dynamics of an electron\nwith mass mand charge\u0000eis governed by the Dirac equation\n[18]\ni@\n@t (x;t)=\u0014\nc\u0012\n\u0000ir+e\ncA(x;t)\u0013\n\u0001\u000b+\fmc2\u0015\n (x;t):(2)\nIn (2), we have introduced the vector \u000b=(\u000b1;\u000b2;\u000b3), where\n\u000biand\fare the Dirac matrices in standard representation [ 19].\nBE\npkpE\n\u0015=2p=pk+pE\n#\nFIG. 1: (Color online.) Schematic setup. An electron with momentum\npis incident at an angle #on a linearly polarized standing laser wave\nwith electric and magnetic components EandB. The momentum p\nhas components pEandpkalong the laser’s electric field Eand wave\nvector k, respectively. The electron is initially spin-polarized along\nthe electric field component. After Kapitza-Dirac scattering, parts of\nthe electron wave packet may have flipped their spin orientation.arXiv:1204.0239v2 [quant-ph] 24 Jul 20122\nThe monochromatic light wave (1) allows us to decompose the\nwave function (x;t) into a discrete set of plane waves, viz.,\n (x;t)=X\nn;\u0010c\u0010\nn(t) \u0010\nn;p(x); \u0010\nn;p(x)=u\u0010\nn;pei(p+nk)\u0001x:(3)\nThe function \u0010\nn;p(x)denotes a free particle Dirac eigenfunc-\ntion of momentum p+nk(n=0;\u00061;\u00062;:::). The index\n\u00102f+\";+#;\u0000\";\u0000#g labels the sign of the energy and the\nspin projection along the laser electric field vector. Taking\nadvantage of the basis functions’ orthonormality the ansatz (3)\nyields\ni˙c\r\nn(t)=ih \r\nn;pj˙ i=\u000f\rE(p+nk)c\r\nn(t)\n\u0000w(t)esin(ckt)\n2kX\n\u0010hu\r\nn;pjE\u0001\u000bju\u0010\nn\u00001;pic\u0010\nn\u00001(t)\n\u0000w(t)esin(ckt)\n2kX\n\u0010hu\r\nn;pjE\u0001\u000bju\u0010\nn+1;pic\u0010\nn+1(t);(4)\nwhere we have introduced the relativistic energy momentum\ndispersion relation E(p)=p\nm2c4+c2p2and the signum \u000f\r,\nwhich is 1 for \r2f+\";+#gand\u00001 for\r2f\u0000\";\u0000#g.\nGeneralized Bragg condition The elastic scattering of\nan electron on a standing light wave may be characterized by a\nBragg condition [ 1] provided that the ponderomotive energy\nof the electron is small (so-called Bragg regime) [5, 20]. This\nBragg condition may be generalized to an inelastic process\nof absorbing and emitting an arbitrary number of photons by\nutilizing momentum conservation p0=p+(nr\u0000nl)kand\nenergy conservation E(p0)=E(p)+(nr+nl)ck;where pand\np0are the initial and final electron momenta. The integers nr\nandnldenote the net numbers of photons exchanged with the\nright- and left-traveling laser waves, respectively, with positive\n(negative) values indicating photon absorption (emission). The\nmomentum and energy conservation laws yield the relativistic\ngeneralized Bragg condition\ncos#\n\u0015p=\n\u0000nr\u0000nl\n2\u0015+nr\u0000nl\njnr\u0000nljnr+nl\n2s\n1\n\u00152\u00001\nnrnl0BBBB@sin2#\n\u00152p+1\n\u00152\nC1CCCCA(5)\nby introducing the angle #(see Fig. 1), the de Broglie wave-\nlength\u0015p=2\u0019=jpj, and the Compton wavelength \u0015C=\n2\u0019=(mc). To be consistent with the nonrelativistic limit nr\nandnlmust have opposite signs. Equation (5) reduces to the\nBragg condition of the two-photon Kapitza-Dirac e \u000bect [1,20]\nby setting nr=\u0000nl=1.\nFrom Eq. (5), it follows that for inelastic processes\n(nr+nl,0)either the initial electron momentum por the\nlaser photon momentum kmust be of the order of mc, i. e., rela-\ntivistic, except we allow for a very large number of interacting\nphotons. Thus, an analysis of inelastic Kapitza-Dirac scattering\ndemands a relativistic treatment by the Dirac equation.\n0 500 1000 1500 2000 2500\nT(laser periods)0.00.20.40.60.81.0probability|c+↑\n0|2\n|c+↑\n3|2\n|c+↓\n3|2FIG. 2: Rabi oscillations of the spin resolved di \u000braction probabilities\nas a function of the interaction time Tfor a three-photon Kapitza-\nDirac e \u000bect. Starting from a pure spin-up state the electron is either\ndi\u000bracted with probability jc+\"\n3(T)j2+jc+#\n3(T)j2or passes the laser\nbeam without momentum transfer with probability jc+\"\n0(T)j2. Laser\nparameters of this simulation correspond to two counterpropagating\nX-ray laser beams with a peak intensity of 2:0\u00021023W=cm2each\nand a photon energy of 3:1keV. The electron impinges at an angle\nof inclination of #=0:4°and a momentum of 176keV=c, in order to\nfulfill the Bragg condition (5).\nA setup for spin-sensitive Kapitza-Dirac scattering\nEquation (4) couples momentum components having momenta\nthat di \u000ber by\u0006kand equal or opposite spin orientation. An\nexplicit calculation of hu\r\nn;pjE\u0001\u000bju\u0010\nn\u00061;pireveals that if pandE\nare orthogonal then c\r\nn(t)couples only to components c\u0010\nn\u00061(t)\nhaving opposite spin orientation. Therefore, a distinct spin\ndynamics may be expected for Kapitza-Dirac scattering with\nan odd number of photons provided that the initial electron\nmomentum pis almost orthogonal to the electric field E. Thus,\nwe focus on a three-photon Kapitza-Dirac e \u000bect, i. e., nr=2,\nnl=\u00001, in the subsequent sections.\nThe condition (5) allows us to determine the initial electron\nmomentum pand the laser wave number kfor a resonant\nthree-photon Kapitza-Dirac e \u000bect. The impinging electron is\nmodeled by a plane wave; thus, the initial condition c+\"\n0(0)=1\nandc\u0010\nn(0)=0else will be applied for the remainder of the\nLetter. We solve the Dirac equation (4) until time Tto compute\nthe transition amplitudes c\u0010\nn(T).\nSpin-flip probability and Rabi frequency Figure 2\nshows the final occupation probabilities jc\u0010\nn(T)j2after laser-\nelectron interaction, which have been calculated by numerical\npropagation of the Dirac equation (4) by a combination of the\nexplicit and the implicit Euler method [ 21]. The laser profile\nwas modeled by an envelope function w(t)that starts with a\nsin2-shaped turn-on ramp of 10 laser cycles and finishes with\nasin2-shaped turn-o \u000bramp of 10 laser cycles having a flat\nplateau in between. The experimental setup was chosen to\nmeet the Bragg condition (5) for a three-photon process. As\nillustrated in Fig. 2, only zeroth-order andthird-order modes\nare nonzero after interaction. The occupation probabilities\noscillate in Rabi cycles of frequency \nR\njc+\"\n0(T)j2=cos2(\nRT=2); (6a)\njc+\"\n3(T)j2+jc+#\n3(T)j2=sin2(\nRT=2); (6b)3\n0.0 0.5 1.0 1.5 2.0\n|pE|/k0.00.20.40.60.81.0Pflip\nFIG. 3: The spin-flip probability Pflipas a function of the electron\nmomentumjpEjin electric field direction. The solid black line is given\nby (8). White squares result from simulations with the Dirac equation\n(4). All simulation parameters except pEare the same as those for\nFig. 2.\nsimilarly to the two-photon Kapitza-Dirac e \u000bect [2] and the\nKapitza-Dirac e \u000bect in atomic beams [ 8]. In the present case,\nhowever, the scattered portion of the electron wave packet\nconsists of two parts which are distinguished by opposite spin\norientations. The period of the Rabi oscillations is 2\u0019=\nR=\n1:9 fs for parameters of Fig. 2.\nFor short times with \nRT\u001c1, the Dirac equation (4) can\nalso be solved analytically via time-dependent perturbation\ntheory. This yields for parameters compatible with the three-\nphoton Bragg condition (5) the probabilities\njc+\"\n3(T)j2= 1\n2\n0T!2 5p\n2jpEj\nk!2\n; (7a)\njc+#\n3(T)j2= 1\n2\n0T!2\n; (7b)\nwith\n0=e3jEj3=(24m3c5k2). ThejEj3dependence clearly in-\ndicates the three-photon nature of the transition. From Eq. (7),\nwe can derive the spin-flip probability Pflipwithin the scattered\nportion of the electron wave packet\nPflip\u0011jc+#\n3(T)j2\njc+\"\n3(T)j2+jc+#\n3(T)j2=1\n25\n2\u0010jpEj\nk\u00112+1: (8)\nThe Rabi frequency may be derived by identifying (7) with the\nTaylor expansion of (6b) for short times T, resulting in\n\nR= \n 0s\n25\n2 jpEj\nk!2\n+1: (9)\nFigure 3 compares the spin-flip probability as a function of\njpEj=k, as obtained from the numerical solution of the Dirac\nequation (4), with our analytical result (8). The initial elec-\ntron momentum in laser propagation direction pkis adjusted\naccording to equation (5) for each value of pE. The analytical\nformula (8) shows very good agreement with the numerical\ndata.Considerations on nonrelativistic theory The spin-\nflip probability (8) can be understood on a qualitative level by\nanalyzing the leading order of the Foldy-Wouthuysen expan-\nsion [ 22] of the Dirac equation (2) which equals the nonrel-\nativistic Pauli equation. This equation features two coupling\nterms which are linear in the fields, namely eA\u0001p=(mc)and\ne\u001b\u0001B=(2mc), where \u001bdenotes the vector of Pauli matrices.\nBoth terms give rise to couplings between adjacent electron\nmomentum components (di \u000bering by\u0006k), with the first term\npreserving and the second term flipping the electron spin. Note\nthat such nearest-neighbor couplings are necessarily involved\nin Kapitza-Dirac scattering with an odd number of photons.\nThe relative strength of the A\u0001pterm as compared with the \u001b\u0001B\nterm is just 2jpEj=k. This results in the spin-flip probability\nPflip, nonrel. =1\n4\u0010jpEj\nk\u00112+1; (10)\nwhich agrees with (8) up to a scale parameter 25=8. The\nprobability (10) can also be derived more rigorously via time-\ndependent perturbation theory for the Pauli equation, which\nyields the nonrelativistic Rabi frequency\n\nR, nonrel. =243\n128\n0s\n4 jpEj\nk!2\n+1: (11)\nNote that the relativistic and nonrelativistic spin-flip proba-\nbilities (8) and (10) and the relativistic and nonrelativistic\nRabi frequencies (9) and (11) agree qualitatively. However,\nthe nonrelativistic expressions cannot be recovered from the\nrelativistic ones by taking a nonrelativistic limit. This is a\nconsequence of the three-photon Bragg condition (5) that en-\nforces a relativistic photon momentum and /or a relativistic\nelectron momentum highlighting the relativistic nature of the\nthree-photon Kapitza-Dirac e \u000bect.\nThe role of the spin Equation (7) indicates that spin-\npreserving transitions in the three-photon Kapitza-Dirac e \u000bect\nare completely suppressed for setups with p?E. This means\nthat under such conditions the scattering is rendered possi-\nble only because the electron does carry spin, which clearly\ndemonstrates the pivotal role the electron spin can play in\nKapitza-Dirac scattering processes.\nThe above argument suggests that for a spinless particle with\np?Ethethree-photon channel of Kapitza-Dirac scattering is\nnot accessible at all. This expectation is confirmed by Fig. 4 (a).\nIt compares numerical results on the Rabi frequency as follow-\ning from the Dirac equation (4) with corresponding numbers\nthat we obtained by solving the time-dependent Klein-Gordon\nequation [ 23]. The predictions significantly di \u000ber in the limit\npE!0. While the Rabi frequency converges to a finite value\nfor spin-half particles [see also (9)], it approaches zero for\nspinless particles, indicating that the scattering channel closes\nindeed. In the spinless case, the spin-flipped channel (7b) is\nmissing, which consequently yields the Rabi frequency\n\nR;spinless = \n 05p\n2jpEj\nk: (12)4\n0.00.20.40.60.81.0\n|pE|/k01234ΩR/Ω0a)\n0 3Klein-Gordon\n0 30.00.20.40.60.81.0probabilityDirac\ndiffraction orderb)\nFIG. 4: Panel (a): The Rabi frequency \nRas a function of the elec-\ntron momentumjpEjin electric field direction for the Dirac equa-\ntion (squares) and the Klein-Gordon equation (triangles). The solid\nblack line is given by (9) and the dashed black line is given by (12).\nPanel (b): The di \u000braction probability after an interaction time of\n0.36 fs for particles with and without spin for parameters as in Fig. 2.\nThe light (dark) gray bars represent the spin-up (spin-down) probabil-\nities. In the case of the Klein-Gordon equation there is no spin degree\nof freedom and, therefore, no dark gray bars appear. Note that the\ndi\u000braction probability depends on the spin degree of freedom.\nFor nonzero values of pE, also Klein-Gordon particles may be\nscattered. But the scattering probability still may be consider-\nably di \u000berent from the Dirac case, as the example in Fig. 4 (b)\nshows.\nExperimental realization An experimental realization\nof the three-photon Kapitza-Dirac e \u000bect may utilize intense\nphoton beams at near-future X-ray laser facilities to form stand-\ning waves. In our numerical simulations, we assumed 3.1 keV\nphotons as envisaged, for example, at the European X-ray free\nelectron laser facility (XFEL) [ 9], which is currently under\nconstruction. The design value of the peak power at this pho-\nton energy is 80 GW. Assuming a focus diameter of 7 nm [ 25],\na field intensity of about 2\u00021023W=cm2results. Laser pulses\nwith duration of about half a Rabi period [which is about 1 fs\nfor the parameters in Fig. 2)] are required for experimental re-\nalization [ 26]. Since the Rabi frequency is much lower than the\nlaser frequency, the photon energy and the electron momentum\nmust be fine-tuned to achieve a resonant transition. Numeri-\ncal simulations indicate that only electrons whose momentum\nvaries by 0:1keV=caround the mean value of 176keV=care\ndi\u000bracted. The photon pulse of the European XFEL with a\nseeded beam of a primary undulator is expected to be coher-\nent, featuring a photon energy uncertainty far below 0.1 keV\n[27]. We note that the electron may lose energy due to sponta-\nneous photoemission with resulting quantitative modification\nof the presented results. Spontaneous emission (scaling with\njEj2), however, is substantially suppressed as compared with\nthe very fast momentum transfer through the three-photon\nKapitza-Dirac e \u000bect which takes place on a femtosecond time\nscale during a Rabi period ( 1=\nR\u00181=jEj3). A numerical so-\nlution of the Landau-Lifshitz equation [ 28] indicates that the\nmomentum transfer into laser propagation direction caused\nby spontaneous emission is su \u000eciently small in order not to\nviolate the resonance condition (5) for the current parameters.\nFinally, the electron beam is di \u000bracted almost in the electron\npropagation direction by 3\u00023:1keV=c. Therefore, a spec-trometer with a resolution below 10keV=cshould be able to\nseparate the di \u000bracted electron beam from the not di \u000bracted\none. In the di \u000bracted beam, about one out of three electrons are\nspin flipped in the case of the scenario in Fig. 2. This spin-flip\nfraction is independent of the interaction time Tof the electron\nwith the laser, in accordance with (8).\nConclusions Pronounced spin e \u000bects in Kapitza-Dirac\nscattering involving three X-ray laser photons interacting with\na weakly relativistic electron beam have been revealed. To this\nend, we deduced a generalized Bragg condition and developed\na theoretical description of the quantum dynamics based on\nthe Dirac equation. The process features characteristic Rabi\noscillations and a competition between spin-preserving and\nspin-flipping nearest-neighbor couplings. The spin-flipping\ntransition becomes dominant in the limit of small angles of\ninclination, where three-photon Kapitza-Dirac scattering cru-\ncially relies on the nonzero electron spin. Our predictions\nmay be tested with the aid of near-future high-intensity XFEL\nsources.\n\u0003heiko.bauke@mpi-hd.mpg.de\nymueller@tp1.uni-duesseldorf.de\n[1]P. L. Kapitza and P. A. M. Dirac, Math. Proc. Cambridge Philos.\nSoc. 29, 297 (1933).\n[2] H. Batelaan, Rev. Mod. Phys. 79, 929 (2007).\n[3]D. L. Freimund, K. Aflatooni, and H. Batelaan, Nature 413, 142\n(2001); D. L. Freimund and H. Batelaan, Phys. Rev. Lett. 89,\n283602 (2002).\n[4]O. Smirnova, D. L. Freimund, H. Batelaan, and M. Ivanov, Phys.\nRev. Lett. 92, 223601 (2004); X. Li, J. Zhang, Z. Xu, P. Fu, D.-S.\nGuo, and R. R. Freeman, Phys. Rev. Lett. 92, 233603 (2004);\nP. Sancho, Phys. Rev. A 82, 033814 (2010).\n[5] M. V . Fedorov, Sov. Phys.–JETP 25, 952 (1967).\n[6]R. Gush and H. P. Gush, Phys. Rev. D 3, 1712 (1971); M. V .\nFedorov, Opt. Commun. 12, 205 (1974).\n[7]P. H. Bucksbaum, D. W. Schumacher, and M. Bashkansky, Phys.\nRev. Lett. 61, 1182 (1988).\n[8]P. L. Gould, G. A. Ru \u000b, and D. E. Pritchard, Phys. Rev. Lett.\n56, 827 (1986); C. Adams, M. Sigel, and J. Mlynek, Phys. Rep.\n240, 143 (1994); M. Freyberger, A. Herkommer, D. Krähmer,\nE. Mayr, and W. Schleich, in Adv. At., Mol., Opt. Phys. , V ol. 41,\nedited by B. Bederson and H. Walther (Academic Press, New\nYork, 1999) pp. 143–180.\n[9]M. Altarelli, R. Brinkmann, M. Chergui, W. Decking, B. Dob-\nson, S. Düsterer, G. Grübel, W. Grae \u000b, H. Graafsma, J. Hajdu\net al. , eds., The European X-Ray Free-Electron Laser Techni-\ncal design report (DESY XFEL Project Group European XFEL\nProject Team Deutsches Elektronen-Synchrotron Member of the\nHelmholtz Association, Hamburg, 2007).\n[10] B. W. J. McNeil and N. R. Thompson, Nature Photon. 4, 814\n(2010); P. Emma, R. Akre, J. Arthur, R. Bionta, C. Bostedt,\nJ. Bozek, A. Brachmann, P. Bucksbaum, R. Co \u000bee, F.-J. Decker\net al. , Nature Photon. 4, 641 (2010); G. Mourou and T. Tajima,\nScience 331, 41 (2011).\n[11] V . M. Haroutunian and H. K. Avetissian, Phys. Lett. A 51, 320\n(1975); M. V . Federov and J. K. McIver, Opt. Commun. 32, 179\n(1980).\n[12] L. Rosenberg, Phys. Rev. A 70, 023401 (2004).5\n[13] D. L. Freimund and H. Batelaan, Laser Phys. 13, 892 (2003).\n[14] P. Krekora, Q. Su, and R. Grobe, J. Phys. B: At., Mol. Opt.\nPhys. 34, 2795 (2001); M. W. Walser, D. J. Urbach, K. Z.\nHatsagortsyan, S. X. Hu, and C. H. Keitel, Phys. Rev. A 65,\n043410 (2002).\n[15] S. X. Hu and C. H. Keitel, Phys. Rev. Lett. 83, 4709 (1999).\n[16] F. H. M. Faisal and S. Bhattacharyya, Phys. Rev. Lett. 93, 053002\n(2004).\n[17] C. Szymanowski, R. Taïeb, and A. Maquet, Laser Phys. 8, 102\n(1998); D. Ivanov, G. Kotkin, and V . Serbo, Eur. Phys. J. C 36,\n127 (2004); F. Ehlotzky, K. Krajewska, and J. Z. Kami ´nski, Rep.\nProg. Phys. 72, 046401 (2009).\n[18] In this article, we use Gauss units and define ~=1.\n[19] B. Thaller, The Dirac Equation , Texts and Monographs in\nPhysics (Springer, 1992).\n[20] H. Batelaan, Contemp. Phys. 41, 369 (2000).\n[21] Other numerical schemes for solving the time-dependent Dirac\nequation have been developed recently in J. W. Braun, Q. Su,\nand R. Grobe, Phys. Rev. A 59, 604 (1999); S. Selstø, E. Lin-\ndroth, and J. Bengtsson, Phys. Rev. A 79, 043418 (2009); G. R.\nMocken and C. H. Keitel, Comp. Phys. Comm. 178, 868 (2008);\nH. Bauke and C. H. Keitel, Comp. Phys. Comm. 182, 2454(2011); F. Fillion-Gourdeau, E. Lorin, and A. D. Bandrauk,\nComp. Phys. Comm. 183, 1403 (2012).\n[22] L. L. Foldy and S. A. Wouthuysen, Phys. Rev. 78, 29 (1950).\n[23] An expansion of the wave function into momentum eigenfunc-\ntions of the Klein-Gordon equation in Hamiltonian form [ 24]\nyields a system of ordinary di \u000berential equations similar to (4).\n[24] H. Feshbach and F. Villars, Rev. Mod. Phys. 30, 24 (1958);\nM. Ruf, H. Bauke, and C. H. Keitel, J. Comp. Phys. 228, 9092\n(2009).\n[25] H. Mimura, S. Handa, T. Kimura, H. Yumoto, D. Yamakawa,\nH. Yokoyama, S. Matsuyama, K. Inagaki, K. Yamamura, Y . Sano\net al. , Nat. Phys. 6, 122 (2010).\n[26] A. A. Zholents and G. Penn, Phys. Rev. ST Accel. Beams 8,\n050704 (2005).\n[27] G. Geloni, V . Kocharyan, and E. Saldin, J. Mod. Opt. 58, 1391\n(2011).\n[28] L. D. Landau and E. Lifshitz, The Classical Theory of Fields ,\n4th ed. (Butterworth-Heinemann, Oxford, 1975); C. H. Keitel,\nC. Szymanowski, P. L. Knight, and A. Maquet, J. Phys. B: At.,\nMol. Opt. Phys. 31, L75 (1998); A. Di Piazza, K. Z. Hatsagort-\nsyan, and C. H. Keitel, Phys. Rev. Lett. 102, 254802 (2009)." }, { "title": "1506.06891v1.Magnetic_Excitations_of_Spin_Nematic_State_in_Frustrated_Ferromagnetic_Chain.pdf", "content": "arXiv:1506.06891v1 [cond-mat.str-el] 23 Jun 2015Journal of thePhysical Society of Japan LETTERS\nMagneticExcitations ofSpin NematicStateinFrustrated Fe rromagneticChain\nHiroakiOnishi\nAdvanced Science ResearchCenter, Japan AtomicEnergy Agen cy, Tokai, Ibaraki, 319-1195, Japan\nBy exploiting density-matrix renormalization group techn iques, we investigate the dynamical spin structure factor o f\naspin-1/2Heisenbergchainwithferromagneticnearest-neighboran dantiferromagneticnext-nearest-neighborexchange\ninteractionsinanappliedmagneticfield.Inafield-induced spinnematicregime,wefindgaplesslongitudinalandgapped\ntransverse spin excitation spectra, in accordance with qua si-long-ranged longitudinal and short-ranged transverse spin\ncorrelations, respectively. The gapless point coincides w ith the dominant longitudinal spin correlation, whereas th e gap\nposition exhibits a characteristic fielddependence contra dicting the dominant transverse spin correlation.\nFrustrated quantummagnets are a class of interacting spin\nsystemsinwhichtherearecompetinginteractionsthatcann ot\nbe satisfied simultaneously. In general, the combined e ffects\nof frustration and quantum fluctuations prevent convention al\nmagneticorder,andweenvisagetheemergenceofnovelspin\nstates. Even when the system is magnetically disordered, we\ntypicallyfind“hidden”orderdescribedbymultiple-spinor der\nparameters, such as dimer, chiral, and multipole ones. In th is\npaper,weparticularlydiscussaquadrupolestate,calleda spin\nnematic (SN) state.1,2)A schematic picture of the SN state is\nthatspinsaligninaspontaneouslychosenaxis,whiletheys till\nfluctuate within the axis and their directions remain unfixed ,\nresembling the directional order of rod-shaped molecules i n\nthenematicliquidcrystal.\nThe occurrence of the SN state has been demonstrated for\nseveral kindsof frustratedquantumspin models.3–13)Among\nthem, a spin-1 /2 chain with ferromagnetic nearest-neighbor\nJ1<0 and antiferromagnetic next-nearest-neighbor J2>0\nexchangeinteractionsina magneticfield h, describedby\nH=J1/summationdisplay\niSi·Si+1+J2/summationdisplay\niSi·Si+2−h/summationdisplay\niSz\ni,(1)\nhas been attracting much currentinterest because of not onl y\nthe novelty of a theoretically predicted SN state,3–9)but also\nthe relevance to a wide variety of edge-shared copper-oxide\nchaincompounds.14–23)\nOnthetheoreticalside,theground-statephasediagramhas\nbeenstudiedindetail,6,7)andit isnowwell establishedthata\nSNstateexistsathighfields,wherethequadrupolecorrelat ion\n/angbracketleftS+\niS+\ni+1S−\njS−\nj+1/angbracketrightexhibits a quasi-long-rangeorder due to the\nformationoftwo-magnonboundstates.Thelongitudinalspi n\ncorrelation/angbracketleftSz\niSz\nj/angbracketrightalsoshowsaquasi-long-rangeorder,while\nthe transverse spin correlation /angbracketleftS+\niS−\nj/angbracketrightdecays exponentially.\nAccordingly, the SN state should have gapless excitations i n\nthe quadrupoleandlongitudinalspin channels,whereasthe re\nshouldbeafiniteexcitationgapinthetransversespinchann el,\ncorresponding to the binding energy of a magnon pair. Note\nthatintheSNregime,thequadrupolecorrelationisdominan t\nbelow the saturation, and the longitudinal spin-density-w ave\n(SDW)correlationbecomesdominantasthemagneticfieldis\ndecreased, although both correlations are quasi-long-ran ged.\nThus,hereafter,we referto this phase as the SN /SDWphase.\nWhenthemagneticfieldisfurtherdecreased,theground-sta te\nphasechangestoa vectorchiral(VC)phaseatlow fields.Concerning the experimentalrealization of the SN state in\nthefruatratedferromagneticchain,LiCuVO 4hasbeenstudied\nfrequently as a prototypical material.14–18)In a field-induced\nphase,neutrondiffractionexperimentshavefounda collinear\nspin-modulatedstructure,whichisconsistentwiththeore tical\nresults for the longitudinal SDW correlation in the SN /SDW\nphase.15,17)Thespincorrelationsexhibitshort-rangebehavior\nin all directions.17)These observations are indeed suggestive\nof the development of the quadrupole correlation. However,\nsince the direct observation of the non-magnetic SN state is\ndifficult,the orderparameterisstill notidentifiedyet.\nTo collect evidence that the SN state occurs in reality, it is\nimportantto find any specific indicationsin what we observe\nby using various microscopic probes. In this context, recen t\ntheoretical studies have pointed out that the NMR relaxatio n\nrate shows characteristic temperature and field dependenci es\nin the SN/SDW phase.24,25)NMR experiments for LiCuVO 4\nhave reported consistent results with theoretical predict ions,\nsignaling the formation of the bound magnon pairs.18)The\npropertiesof thelow-energyspin excitationspectrahavea lso\nbeendiscussed.24,26–28)\nIn this paper, to clarify the propertyof the SN /SDW phase\nfrom the viewpoint of the spin dynamics,we investigate spin\nexcitation spectra in a wide range of momentum and energy\nby numerical methods. We clearly find that the longitudinal\nspin excitation spectrum is gapless, while the transverse o ne\nis gapped, as naively expected from the behavior of the spin\ncorrelations. We discuss the field dependence of the spectra l\nweighttransferinrelationwiththedominantspincorrelat ion.\nA striking feature is that the momentum position of the gap\ndeviatesfromthatofthedominanttransversespincorrelat ion\nasthesystem approachesthesaturation.\nLet us consider the model (1) in an N-site chain, and take\nJ2=1 as the energy unit. We investigate the spin excitation\ndynamics at zero temperature by exploiting density-matrix\nrenormalizationgroup (DMRG) techniques.29,30)We employ\nthe finite-system algorithm in open boundaryconditions. We\ncomputethedynamicalspinstructurefactor,definedby\nSα(q,ω)=−1\nπIm/angbracketleftψG|Sα†\nq1\nω+EG−H+iηSα\nq|ψG/angbracketright,(2)\nwhere|ψG/angbracketrightisthegroundstatewitheigenenergy EG.Notethat\n|ψG/angbracketrightis given by the lowest-energy state in the subspace of a\ngiven magnetization m=M/N, whereM=/summationtext\niSz\ni. For the\n1J.Phys. Soc. Jpn. LETTERS\nFig. 1. (Color online) Intensity plots ofthedynamical spinstruct ure factor Sα(q,ω)atJ1=−1andJ2=1,obtained bydynamical DMRGcalculations with\nN=128. (a)Sz(q,ω) atm=0andh=0. (b)Sz(q,ω) and (c)Sx(q,ω) atm=0.125(=16/128) and h=0.649.\ncalculation of Sα(q,ω) atm, we set the magnetic field to be\nthe midpoint of the magnetization plateau of Min theN-site\nsystem.ηis a small broadening factor, and we set η=0.1\nunlessotherwisespecified.Ourparticularinterestistocl arify\nthe anisotropy between longitudinal Sz(q,ω) and transverse\nSx(q,ω)duetotheformationofthetwo-magnonboundstates\ninthefield-inducedSN /SDWphase.\nFor the analysis of the spin excitation spectrum, we use a\ndynamicalDMRGmethod,30)targetingthegroundstate |ψG/angbracketright,\nan excitated state Sα\nq|ψG/angbracketright, and the so-called correction vector\n[ω+EG−H+iη]−1Sα\nq|ψG/angbracketright.Here,wenotethatthetruncation\nerrorrapidlyincreasesasthenumberoftargetstatesincre ases.\nTherefore,to obtain Sx(q,ω) with keepinghighaccuracy,we\ncalculate S+(q,ω) andS−(q,ω) separately, and then use the\nrelationSx(q,ω)=[S+(q,ω)+S−(q,ω)]/4,insteadofdirectly\ncalculating Sx(q,ω).Notealsothatwecalculatethespectrum\natqandωafter one DMRG run with fixed qandω, so that\nwe need to perform a great number of DMRG runs to obtain\na fullspectrum.\nLet us start with a brief discussion on the spectrum at zero\nfield. In Fig. 1(a), we present the intensity plot of Sz(q,ω) at\nJ1=−1,J2=1 (energy unit throughout the paper), m=0,\nandh=0. Note that Sz(q,ω)=Sx(q,ω) ath=0 due to the\nSU(2)spinrotationsymmetry.Wefindasinusoidaldispersio n\nthatgivesthelowerboundaryofacontinuum.Thesinusoidal\ndispersion represents the spinon excitation, described by the\ndes Cloizeaux-Pearson mode,31)since the system decouples\ninto two antiferromagnetic chains if J1→0. We see a large\namountof spectral weight at a lowest-energypeak ( q0,ω0)=\n(π/2,0.04). Note here that the spectrum is asymmetric with\nrespectto q0,aspointedoutbythepreviousstudies.14,32)That\nis, the spectralweight mainlylies nearthe lowerboundaryo f\nthe continuum for qq0. On the other hand, ω0coincides\nwitha spinexcitationenergy,\n∆(N,M)=[E0(N,M+1)+E0(N,M−1)−2E0(N,M)]/2,(3)\nwhereE0(N,M) is the lowest energy of the N-site system in\nthe subspace of Math=0. We find that∆(N,0) shifts to\nlower energy as Nincreases, and it is extrapolated to almost\nzero in the limit of N→∞. Note that an exponentiallysmall\ngap has been predicted by renormalization group theory, but\nit ishardto detectsuchasmall gapnumerically.33)\nNow, let us look into the longitudinal and transverse spinexcitation spectra in the SN /SDW phase. In Figs. 1(b) and\n1(c),weshow Sz(q,ω)andSx(q,ω),respectively,at J1=−1,\nm=0.125,andh=0.649.ForSz(q,ω),alowest-energypeak\nisat(q0,ω0)=(0.375π,0.00).Thatis, q0movestowardsmall\nmomentum from the position at zero field. ω0is nearly zero,\nindicatingagaplessmodeforthelongitudinalspinexcitat ion.\nWe see that a certainamountof spectralweight is transferre d\nto the origin due to a finite uniform magnetization, which is\nalso indicative of a gapless mode. These gapless points are\nclearly visible due to the large intensity. Moreover, Sz(q,ω)\nseems to be gapless at q=π−q0andq=π, although we\ndo not observe significant intensity near the possible gaple ss\npointsinq>π/2.\nIn contrast, for Sx(q,ω), we observe a lowest-energypeak\nat(q0,ω0)=(31π/64,0.15).Thatis, q0remainsnear π/2even\nin the SN/SDW phase. On the other hand, ω0appearsto be a\nfiniteenergy.Infact, ω0agreeswiththespinexcitationenergy\n∆(N,M), andit is extrapolatedto a finite valuein the limit of\nN→∞withm=M/Nfixed, indicating a gapped mode for\nthe transverse spin excitation. Note that Sx(q,ω) consists of\nS−(q,ω) andS+(q,ω). As shown in Fig. 2, both spectra have\nthelowest-energypeakatthesameposition,whiletheovera ll\nstructure is different between them. We find that S−(q,ω) is\nhighly dispersive in a wide range of momentum and energy,\nandS+(q,ω) is less dispersive and it is mainly concentrated\nin asmall regionnearthelowest-energypeak.\nHere, let us discuss how the lowest-energy peak position\ndepends on min the whole range of m. In Fig. 3(a), we plot\nq0ofSz(q,ω)asafunctionof mforseveralvaluesof J1.Note\nthat we also find a sharp peak at the origin for finite m, but\nFig. 2. (Coloronline)Intensityplotsofthedynamicslspinstruct urefactors\n(a)S−(q,ω) and (b)S+(q,ω) atJ1=−1,J2=1,m=0.125, andh=0.649.\nNote that Sx(q,ω)=[S+(q,ω)+S−(q,ω)]/4 is plotted in Fig. 1(c).\n2J.Phys. Soc. Jpn. LETTERS\nFig. 3. (Color online) The lowest-energy peak position q0of (a)Sz(q,ω)\nand (b)S−(q,ω) for several values of J1atJ2=1 as a function of m. Solid\nsymbols denote the VC phase at low fields, and open symbols den ote the\nSN/SDW phase at high fields. Note that q0shows a stepwise change simply\ndue to finite-size e ffects. The resolution of the momentum is 2 π/Nand we\nuseN=128 in the present calculations. Here, the broadning factor is set to\nη=0.02 to determine the position of thelowest-energy peak preci sely.\nwefocusonthefield-inducedshiftofthepeakpositionwhich\noriginally locates near π/2 at zero field. At small m, where\nthe system is in the VC phase, q0shows little dependenceon\nm. At large m, where the system is in the SN /SDW phase, q0\nfollows the relation q0=(1/2−m)πregardless of J1, which\nsupportsthebosonizationresult.24)Notethat q0isrepresented\nbythedensityofboundmagnons1 /2−m.\nIn Fig. 3(b), we show the mdependenceof q0ofS−(q,ω).\nInthe VC phase at small m, we noticethat q0agreeswith the\nincommensurabilityofthe transversespincorrelation,as will\nbeshowninFig.4(d).Noteherethatithasbeenrevealedthat\ntheincommensuratewavenumberintheVCphaseisstrongly\nquantum renormalized toward π/2 compared with the pitch\nanglecos−1(−J1/4J2) ofthehelicalorderin theclassical spin\ncase.6)Indeed,weobservethat q0=π/2atJ1=−0.5and−1,\nalthoughthe correspondingclassical pitch angle is 0 .46πand\n0.42π, respectively. For small |J1|, we see that q0remains at\nπ/2below a thresholdvalue of mevenin the SN/SDW phase\natlargem.However, q0goesawayfrom π/2asmapproaches\nthesaturation,andthedeviationfrom π/2ismorepronounced\nfor larger|J1|. On the other hand,the exact energydispersion\noftheone-magnonexcitedstate inthe fullypolarizedstate is\nǫ1(q)=J1(cosq−1)+J2(cos2q−1)+h,(4)\nand its minimum is at q=cos−1(−J1/4J2), which coincides\nwiththe classical pitchangle.4)Thepresentnumericalresults\nat the saturation totally agree with this exact description . We\nmentionthat the bosonizationanalysis showsthat the botto m\noftheone-magnonbandisfoundat π/2,24)butitisvalidonly\nfor the weak-coupling regime |J1|≪J2and inapplicable in\nthe limit of m→1/2. We thus confirm that the bottom of\nthe one-magnonbandmovesfromthe quantumrenormalized\nincommensurate wave number to the classical pitch angle as\nmincreasesfromzerotothesaturation.Thisfeaturewouldbe\nusefultodetermineexchangecouplingsofrealmaterials.\nTo gain an insight into the properties of the field-induced\nspectralweighttransferandthedominantspincorrelation ,we\nexaminethespinstructurefactor,\nSα(q)=/angbracketleftψG|Sα†\nqSα\nq|ψG/angbracketright, (5)\nwith attention to sum rules on the integrated intensity of th e\nFig. 4. (Color online) (a) Sz(q) atJ1=−1, (b)Sx(q) atJ1=−1, and (c)\nSx(q) atJ1=−2.5 for several values of m. (d) Pitch angle Q, determined\nfrom the maximum point of S−(q), as a function of m. Solid symbols denote\nthe VC phase at low fields, and open symbols denote the SN /SDW phase at\nhigh fields. J2=1 is fixed. Data are obtained by DMRG calculations with\nN=128.\ndynamical spin structure factor. To obtain Sα(q), we perform\nground-state DMRG calculations independent of dynamical\nDMRGrunsfor Sα(q,ω),sincewecanobtainaccurateresults\nwithrelativelysmallcomputationalcost.Thesumrulefort he\nenergy-integratedintensityreads\nIα(q)≡/integraldisplaydω\n2πSα(q,ω)=Sα(q), (6)\nleading to the spin structure factor. As for the total intens ity\nIα≡/summationtext\nqSα(q),we have\nIz=Ix=N/4,I±=N/2∓M. (7)\nInFig.4(a),weshow Sz(q)forseveralvaluesof matJ1=−1.\nThere is a clear peak at q=π/2 form=0, and it shifts\ntoward small momentum with increasing m, as indicated by\nvertical arrows. The peak position of Sz(q) representing the\nSDWcorrelationisinagreementwiththelowest-energypeak\npositionq0ofSz(q,ω) in Fig. 3(a). We also find that Sz(q)\ngrows at q=0 asmincreases. In other momentum parts,\nSz(q)is suppressedso as to keep the total intensity Iz=N/4.\nAccordingly, Sz(q,ω)exhibitsthespectralweighttransfer,as\nseenin Fig.1(b).\nWe show Sx(q)atJ1=−1andJ1=−2.5in Figs. 4(b)and\n4(c), respectively. We find a sharp peak structure in the VC\nphase at small m, while the peak structure becomes broad in\nthe SN/SDW phase at large m. This is because the transverse\nspin correlationexhibits an algebraic decay in the VC phase ,\nand it is short-rangedin the SN /SDW phase. As mincreases,\nthe borad peak is leveled o ff, while the baseline goes up, and\neventuallywe finda completelyflat profile,i.e., Sx(q)=1/4,\natthesaturation.Ontheotherhand,thepeakpositiondepen ds\n3J.Phys. Soc. Jpn. LETTERS\nFig. 5. (Color online) The extrapolated gap ∆(m) for several values of J1\natJ2=1 as a function of m. Here we plot data in the SN /SDW phase. Note\nthat the zero-field gap is tiny,33)and the VC phase is supposed to begapless.\nonJ1andm. Indeed, we find a peak near q=π/2 regardless\nofmforJ1=−1 [Fig. 4(b)], while a peak is at q=25π/64\natm=0 and it moves to q=π/2 near the saturation for\nJ1=−2.5[Fig.4(c)].\nTo clarify the mdependenceof the dominant qcomponent\nof the transverse spin correlation,we determinea pitch ang le\nQfrom the maximum point of S−(q), as shown in Fig. 4(d).\nNote that S−(q) andS+(q) have a peak at the same position.\nIn the VC phase at small m,Qagrees with the lowest-energy\npeak position q0ofS−(q,ω), as denoted by solid symbols in\nFigs. 3(b) and 4(d). However, in the SN /SDW phase at large\nm,Qbehaves quite differently from q0asmapproaches the\nsaturation. We find that Qchanges toward π/2 regardless of\nJ1, rather thanthe classical pitchangle cos−1(−J1/4J2) asq0,\nindicatingthatthedominant qcomponentisnotequivalentto\nthemomentumpositionoftheopeninggap.Thisdiscrepancy\nis naturally understood in terms of the short-range nature o f\nthetransversespincorrelation.Itisinadequatetotaketh eonly\nleadingasymptotictermthatrepresentstheexponentialde cay\nofthetransversespincorrelationfunctioninordertodesc ribe\nthe excitation dynamics correctly. In fact, we have seen tha t\nSx(q)turnsinto theflat profileat thesaturation,meaningthat\nnotonlythedominant qcomponentbutalsoothercomponents\nequallycontributetothespectralweight.\nFinally,wepresentthedependenceoftheenergygapofthe\ntransversespin excitationin Fig. 5. We extrapolatefinite- size\ndatatothethermodynamiclimit N→∞byassumingalinear\nrelation∆(N,M)=∆(m)+a/N. Form/greaterorsimilar0.1, we find that\n∆(m) increases as|J1|becomes large for J1=−0.5,−1, and\n−1.5, indicating that the ferromagnetic exchange interaction\nstabilizesthetwo-magnonboundstate. ∆(m)turnstodecrease\nwith further increasing |J1|, shown for J1=−2. For small m,\n∆(m)israpidlyreducedas mdecreasesdowntotheboundary\nbetweentheSN/SDWandVCphases.\nIn summary,we have studied the spin excitation dynamics\nof the frustratedferromagneticchain in the magneticfield b y\nnumerical methods. In the field-induced SN regime, the spin\nexcitationspectraexhibithighlyanisotropicbehaviorbe tween\ngaplesslongitudinalandgappedtransversecompoments.Th e\nfielddependenceofthegaplesspointof Sz(q,ω)isconsistent\nwith the dominant longitudinal spin correlation. In contra st,\nthe gappositionof Sx(q,ω)showsa uniquefield dependence\nthatcontradictswiththedominanttransversespincorrela tion.\nWe hope that these features could be examined by inelasticneutronscatteringexperimentswhensearchingforsignatu res\nof theSN/SDWphase.On theotherhand,we shouldobserve\ngaplessexcitationsin the quadrupolechannel.It wouldbe a n\ninteresting future problem to study the quadrupoleexcitat ion\ndynamicsanditspossiblerelevancetoobservables.34)\nAcknowledgment The author thanks T. Masuda, S. Maekawa, M. Mori,\nand T. Sugimoto for discussions and comments. Part of comput ations were\ncarried out on the supercomputers at the Japan Atomic Energy Agency and\nthe Institute for Solid State Physics,the University of Tok yo.\n1) A.F.Andreev and I. A.Grishchuk, Sov.Phys.JETP 60, 267 (1984).\n2) For a recent review, see K. Penc and A. M. L¨ auchli, in Introduction to\nFrustratedMagnetism ,ed.C.Lacroix,P.Mendels,andF.Mila(Springer-\nVerlag, Berlin, 2011), p. 331.\n3) A.V. Chubukov, Phys.Rev. B 44,4693 (1991).\n4) L. Kecke, T. Momoi, and A. Furusaki, Phys. Rev. B 76, 060407(R)\n(2007).\n5) T. Vekua, A. Honecker, H.-J. Mikeska, and F. Heidrich-Mei sner, Phys.\nRev. B76, 174420 (2007).\n6) T. Hikihara, L. Kecke, T. Momoi, and A. Furusaki, Phys. Rev . B78,\n144404 (2008).\n7) J. Sudan, A. L¨ uscher, and A. M. L¨ auchli, Phys. Rev. B 80, 140402(R)\n(2009).\n8) M.E.ZhitomirskyandH.Tsunetsugu,Europhys.Lett. 92,37001(2010).\n9) M.Sato,T.Hikihara,andT.Momoi,Phys.Rev.Lett. 110,077206(2013).\n10) N. Shannon, T. Momoi, and P. Sindzingre, Phys. Rev. Lett. 96, 027213\n(2006).\n11) H.Tsunetsugu and M.Arikawa, J.Phys.Soc. Jpn. 75,083701 (2006).\n12) A.L¨ auchli, F.Mila, and K. Penc, Phys.Rev. Lett. 97, 087205 (2006).\n13) T. Momoi, P. Sindzingre, and K. Kubo, Phys. Rev. Lett. 108, 057206\n(2012).\n14) M. Enderle, B. Fåk, H.-J. Mikeska, R. K. Kremer, A. Prokofi ev, and W.\nAssmus,Phys.Rev. Lett. 104, 237207 (2010).\n15) T.Masuda,M.Hagihala, Y.Kondoh,K.Kaneko,andN.Metok i, J.Phys.\nSoc. Jpn. 80, 113705 (2011).\n16) L. E. Svistov, T. Fujita, H. Yamaguchi, S. Kimura, K. Omur a, A.\nProkofiev, A. I. Smirnov, Z. Honda, and M. Hagiwara, JETP Lett .93,\n21 (2011).\n17) M. Mourigal, M. Enderle, B. Fåk, R. K. Kremer, J. M. Law, A. Schnei-\ndewind,A.Hiess,andA.Prokofiev,Phys.Rev.Lett. 109,027203(2012).\n18) K.Nawa,M.Takigawa,M.Yoshida,andK.Yoshimura,J.Phy s.Soc.Jpn.\n82, 094709 (2013).\n19) M.Hase, H.Kuroe, K.Ozawa, O.Suzuki, H.Kitazawa, G.Kid o, and T.\nSekine, Phys.Rev. B 70, 104426 (2004).\n20) A. A. Bush, V. N. Glazkov, M. Hagiwara, T. Kashiwagi, S. Ki mura, K.\nOmura,L.A.Prozorova, L.E.Svistov, A.M.Vasiliev, and A.Z heludev,\nPhys.Rev. B 85, 054421 (2012).\n21) B.Willenberg, M.Sch¨ apers, K.C.Rule, S.S¨ ullow,M.Re ehuis, H.Ryll,\nB. Klemke, K.Kiefer, W. Schottenhamel, B. B¨ uchner, B. Oula ddiaf, M.\nUhlarz, R.Beyer, J.Wosnitza, and A.U.B.Wolter, Phys.Rev. Lett.108,\n117202 (2012).\n22) S.E.Dutton, M.Kumar, M.Mourigal, Z.G.Soos,J.-J.Wen, C.L.Bro-\nholm, N. H. Andersen, Q. Huang, M. Zbiri, R. Toft-Petersen, a nd R. J.\nCava, Phys.Rev. Lett. 108, 187206 (2012).\n23) K.Nawa,Y.Okamoto,A.Matsuo,K.Kindo,Y.Kitahara, S.Y oshida, S.\nIkeda, S.Hara,T.Sakurai, S.Okubo,H.Ohta,and Z.Hiroi,J. Phys.Soc.\nJpn.83,103702 (2014).\n24) M.Sato,T.Momoi,andA.Furusaki,Phys.Rev.B 79,060406(R)(2009).\n25) M.Sato, T.Hikihara, and T.Momoi,Phys.Rev. B 83,064405 (2011).\n26) O.A. Starykh and L.Balents, Phys.Rev. B 89,104407 (2014).\n27) H.Onishi, J.Phys.:Conf. Ser. 592, 012109 (2015).\n28) A. Smerald, H. T. Ueda, and N. Shannon, Phys. Rev. B 91, 174402\n(2015).\n29) S.R. White, Phys.Rev. Lett. 69, 2863 (1992).\n30) E.Jeckelmann, Phys.Rev. B 66, 045114 (2002).\n31) J.des Cloizeaux and J.J.Pearson, Phys.Rev. 128, 2131 (1962).\n32) J.Ren and J.Sirker, Phys.Rev. B 85,140410(R) (2012).\n33) C. Itoi and S.Qin, Phys.Rev. B 63, 224423 (2001).\n34) A.V¨ oll and S.Wessel,Phys.Rev. B 91,165128 (2015).\n4" }, { "title": "1405.4164v1.Inferring_hidden_states_in_a_random_kinetic_Ising_model__replica_analysis.pdf", "content": "arXiv:1405.4164v1 [cond-mat.dis-nn] 16 May 2014Inferring hidden states in a random kinetic Ising\nmodel: replica analysis\nLudovica Bachschmid Romano and Manfred Opper\nDepartment of Artificial Intelligence, Technische Universit¨ at Ber lin, Marchstraße 23,\nBerlin 10587, Germany\nE-mail:ludovica.bachschmidromano@tu-berlin.de and\nmanfred.opper@tu-berlin.de\nAbstract. We consider the problem of predicting the spin states in a kinetic Ising\nmodel when spin trajectories are observed for only a finite fractio n of sites. In a\nBayesian setting, where the probabilistic model of the spin dynamics is assumed to\nbe known, the optimal prediction can be computed from the conditio nal (posterior)\ndistribution ofunobserved spins given the observedones. Using th e replica method, we\ncompute the error of the Bayes optimal predictor for parallel disc rete time dynamics in\nafullyconnectedspinsystemwithnonsymmetricrandomcouplings. T heresults, exact\nin the thermodynamic limit, agree very well with simulations of finite spin systems.\n1. Introduction\nThe problem of statistical inference in kinetic Ising models has recen tly attracted\nconsiderable interest in the statistical physics community, see e.g. [1–5]. These systems\ncan be viewed as simple models of networks of spiking neurons and pro vide a prototype\nmodel for which a reconstruction of the network from dynamical d ata can be studied.\nBased on a temporal sequence of observed spin variables, a major goal is to estimate\nthe couplings between sites. This task gets more complicated when a t some sites the\nspin trajectories are not observed. Besides the problem of inferr ing the couplings\nit is then also interesting to predict the states of the non observed spins when the\ncouplings are known. In fact, an iterative solution to the maximum like lihood problem\nfor estimating the couplings is the Expectation Maximization (EM) algorithm [6] which\nwould iterate between estimating hidden spin states (given the last e stimate of the\ncouplings) and reestimating the couplings. Unfortunately, exact in ference of hidden\nstates is not tractable for large networks, but algorithms which ar e based on statistical\nphysics approximations have recently been discussed [7,8]. Hence, it will be interesting\nand important to study a scenario for which the theoretically optima l performance for\npredicting hidden spins can be computed exactly. In this paper, we w ill show that\nsuch a solution can be found in the thermodynamic limit of an infinitely lar ge network\nwhen the couplings are random. Our approach will be based on the re plica method of\ndisordered systems which enables us to compute quenched averag es over the randomInferring hidden states in a random kinetic Ising model: rep lica analysis 2\ncouplings for thermodynamic quantities of the model. These thermo dynamic quantities\nare themselves functions of posterior averages (e.g. local magne tizations) of the hidden\nspins. The replica approach has been successfully applied in the past to a large variety\nof statistical learning problems for staticnetwork models (for a summary see [9–11]).\nWe will restrict ourselves to a model where the couplings are mutually independent\nrandom variables, i.e. where no symmetry between in-and outgoing c onnections are\nassumed. For such type of models (without the observations) var ious exact solutions\nfor the non equilibrium dynamics have been computed, see e.g. [1,5] a nd [12,13] for\nsoft spin models. From the point of view of equilibrium statistical phys ics the case\nof symmetric couplings might be interesting. Such a spin model would o bey detailed\nbalance and allow for a stationary Gibbs distribution. Unfortunately , for the Ising case,\nthe exact computation of time dependent correlation functions wh ich are necessary for\nour analysis seems not possible. On the other hand, from a point of v iew of neural\nmodeling, the assumption of symmetric couplings is not realistic [1,14], as synaptic\nconnections in biological networks are known to be strongly asymme tric. Hence, we\nbelieve that our restriction to asymmetric couplings is justified both from a modeling\nand a computational perspective.\n2. The model and Bayes optimal inference\nWe will consider a model with NIsing spins which are divided into two groups: a group\nof spinssi(t) at sites i=1,..., Nobs=λNwhich are observed during a time interval of\nTtime steps, and a group of hidden, i.e. unobserved spins, denoted b yσa(t) at sites\na= 1,...,N hid= (1−λ)N. We assume parallel Markovian dynamics for the entire spin\nsystem, which is governed by the transition probability\nP[{s,σ}(t+1)|{s,σ}(t)] =/productdisplay\niesi(t+1)gi(t)\n2cosh[gi(t)]/productdisplay\naeσa(t+1)ga(t)\n2cosh[ga(t)], (1)\nwhere the fields are defined as\ngi(t) =/summationdisplay\njJijsj(t)+/summationdisplay\nbJibσb(t), g a(t) =/summationdisplay\njJajsj(t)+/summationdisplay\nbJabσb(t),(2)\nin terms of the couplings Jand{s,σ}denotes all the possible spin vector configurations;\nwhen the time index is not specified we are considering the whole time se ries,t= 0...T.\nThe total probability for a spin trajectory is given by\nP({s,σ}) =1\n2NT−1/productdisplay\nt=0P[{s,σ}(t+1)|{s,σ}(t)], (3)\nwhere we have considered completely random initial condition P0[{s,σ}(0)] = 1/2N.\nTo make predictions on the unobserved spins σa(t), we assume that the model given\nby the couplings Jis perfectly known and the posterior, i.e. conditional probability ofInferring hidden states in a random kinetic Ising model: rep lica analysis 3\nthe hidden spins defined by\nP({σ}|{s}) =P({s,σ})\nP({s}), (4)\ngives the complete information for an optimal inference of hidden sp ins. Based on this\nprobabilistic information, the best possible prediction σopt\na(t) for the hidden spin at site\naand at time tis computed by\nσopt\na(t) = sign[ma(t)], (5)\nwhere the local magnetization is defined as the posterior expectat ion\nma(t) =/summationdisplay\n{σ}σa(t)P({σ}|{s}). (6)\nNote that this does not correspond to the most likely spin configuration {σ}, because\nwe have averaged out the configurations of spins σb(t′) forb∝ne}ationslash=aandt′∝ne}ationslash=t.\nGiven a true ‘teacher’ sequence {σ∗}of unobserved spins, we are interested in the\ntotal quality of the Bayes optimal prediction, i.e. in the expected pr obability of wrongly\npredicting a spin at site aand timet, given by the Bayes error\nε=/summationdisplay\n{s,σ∗}P({s,σ∗})Θ(−σ∗\na(t)ma(t)) =/summationdisplay\n{s}P({s})/summationdisplay\n{σ∗}P({σ∗}|{s})Θ(−σ∗\na(t)ma(t)),\n(7)\nwhere the step function Θ( x) = 1 forx>0 and 0 else. In the next section we will use\nthe replica method to compute the error in the thermodynamic limit N→ ∞, when the\ncouplingsJare assumed to be mutually independent Gaussian random variables, with\nzero mean and variance of the order 1 /N.\n3. Replica analysis\nThe posterior statistics of the hidden spins can be obtained from th e following partition\nfunction\nP({s}) =1\n2N/summationdisplay\n{σ}/productdisplay\ntP[{s,σ}(t+1)|{s,σ}(t)], (8)\nwhich equals the total probability of the observed spin configuratio ns and is also the\nnormalizer of the posterior probability. Typical performance in the thermodynamic\nlimit for random couplings are then computed from the quenched ave rage of the free\nenergyF=−∝an}bracketle{tlnP({s})∝an}bracketri}htJ,s, where the average is taken over the the couplings Jand\nover the observed spin configurations with their weights P({s}). Hence, the averaged\nfree energy is given by\nF=−/summationdisplay\n{s}∝an}bracketle{tP({s})logP({s})∝an}bracketri}htJ. (9)Inferring hidden states in a random kinetic Ising model: rep lica analysis 4\nThis average can be computed by the replica trick [9–11] in the follow ing way:\nF=−lim\nn→1d\ndnlog/summationdisplay\n{s}∝an}bracketle{tPn({s})∝an}bracketri}ht. (10)\nFor integer n, we have\n/summationdisplay\n{s}∝an}bracketle{tPn({s})∝an}bracketri}htJ=1\n2nN/summationdisplay\n{s}/summationdisplay\n{σ(1)}.../summationdisplay\n{σ(n)}/angbracketleftBigg/bracketleftBiggn/productdisplay\nα=1exp/braceleftBigg/summationdisplay\nitsi(t+1)gα\ni(t)\n+/summationdisplay\natσα\na(t+1)gα\na(t)−/summationdisplay\nitlog2cosh[gα\ni(t)]−/summationdisplay\natlog2cosh[gα\na(t)]/bracerightBigg/bracketrightBigg/angbracketrightBigg\nJ,(11)\nwith\ngα\ni(t) =/summationdisplay\njJijsj(t)+/summationdisplay\nbJibσα\nb(t), gα\na(t) =/summationdisplay\njJajsj(t)+/summationdisplay\nbJabσα\nb(t).(12)\nTo perform the average over the couplings Jij,Jib,JajandJab, which are assumed to be\nmutually independent Gaussian random variables with zero mean and v ariancek2/N,\nwe note that the fields gα\ni(t) andgα\na(t) are also Gaussian, which are independent for\ndifferent sites ianda, but will be dependent for different replica index αandβand also\npossibly for different times. This yields\n/angbracketleftBig\ngα\ni(t)gβ\ni(t′)/angbracketrightBig\n=/angbracketleftbig\ngα\na(t)gβ\na(t′)/angbracketrightbig\n=k2/parenleftbig\nλS(t,t′)+(1−λ)Qαβ(t,t′)/parenrightbig\n,\n∝an}bracketle{tgα\ni(t)gα\ni(t′)∝an}bracketri}ht=∝an}bracketle{tgα\na(t)gα\na(t′)∝an}bracketri}ht=k2(λS(t,t′)+(1−λ)Cα(t,t′)),(13)\nwhere we have defined the following order parameters\nCα(t,t′) =1\nNhid/summationdisplay\naσα\na(t)σα\na(t′) fort