title
stringlengths 31
247
| content
stringlengths 15
7.12M
|
---|---|
2311.14719v1.Thermal_Spin_Orbit_Torque_with_Dresselhaus_Spin_Orbit_Coupling.pdf | ThermalSpin-OrbitTorquewithDresselhaus
Spin-OrbitCoupling
Chun-YiXue,Ya-RuWang,Zheng-ChuanWang*
SchoolofPhysicalSciences,
UniversityofChineseAcademyofSciences,Beijing100049,China.
*wangzc@ucas.ac.cn
Abstract
BasedonthespinorBoltzmannequation,weobtainatemperature
dependentthermalspin-orbittorqueintermsofthelocalequilibrium
distributionfunctioninatwo-dimensionalferromagnetwith
Dresselhausspin-orbitcoupling.Wealsoderivethecontinuity
equationofspinaccumulationandspincurrent—thespindiffusion
equationinDresselhausferromagnet,whichcontainsthethermal
spin-orbittorqueunderlocalequilibriumassumption.This
temperaturedependentthermalspin-orbittorqueoriginatesfromthe
temperaturegradientappliedtothesystem,itisalsosensitiveto
temperatureduetothelocalequilibriumdistributionfunctiontherein.
Inthespindiffusionequation,wecansingleouttheusualspin-orbit
torqueaswellasthespintransfertorque,whichisconcededtoour
previousresults.Finally,weillustratethembyanexampleof
spin-polarizedtransportthroughaferromagnetwithDresselhaus
spin-orbitcouplingdrivenbytemperaturegradient,thosetorquesincludingthermalspin-orbittorquearedemonstratednumerically.
PACS:75.60.Jk,72.15Jf,75.70.Tj,85.70.-w
I.Introduction
Asanewbranchofspintronics,spin-orbitronicshasbeen
exploredwhateverintheoriesorexperiments[1-2],becauseitcan
provideanefficientwaytomanipulatethelocalmagneticmomentin
thedevicesofspintronicsviaspin-orbittorques[3-5].Sincethe
discoveryofgiantmagnetoresistance[6]andspintransfertorque[7],
nowadaysmagnetoresistancerandomaccessmemory(MRAM)
drivenbyspin-polarizedcurrenthasbeendesignedandrealized
industrially[8-10].ThefirstgenerationMRAMistoggleMRAM
[11],whichconsistofatransistorandamagnetictunneljunction
(MTJ),butthisstructurebringsobviousdisadvantagestothetoggle
MRAMduetothebigmagneticfield[12].Thesecondgeneration
MRAMarespintransfertorqueMRAM(STT-MRAM)[13]and
perpendicularspintransfertorqueMRAM(pSTT-MRAM)[14],in
whichmagnetizationreversalinSTT-MRAMandpSTT-MRAMrely
onthespin-polarizedelectricalcurrentratherthanbigmagneticfield,
sotheyhavefasterwritingspeed.However,STT-MRAMdepends
onthermalactivationtostartswitching[15],soithasaninitial
latencywhichrestrictsitsmaximumcachespeed.Thus,oneproposedthethirdgenerationMRAMtosolvethisproblem,which
isdrivenbyspinorbittorque(SOT).SOTcanbeemployedtoswitch
themagnetizationinasystemwithabrokeninversionsymmetry.
SinceSOTneedalowercriticalcurrent,ithasbetterthermal
stability[16].Recently,ThermalSOThasbeenobservedin
experiments[17-18].Tillnow,SOT-MRAMisagoodcandidateof
magneticstoragedeviceforbetterperformance.
In2007,HatamiproposedSTTwhichcanbedrivenbythermal
spincurrent[19].Similarly,SOTcanalsobedrivenbythermalspin
currentintheprocessofspin-polarizedelectrontransport,whichis
calledthermalSOT(TSOT).TSOTwasfirstlyproposedby
Freimuthin2014intermsofBerryphase[20-21]whichisexpressed
bythequantumstatesofelectrons,whilethestatesusuallyshouldbe
calculatedbytheDensityFunctionTheory(DFT),itissomewhat
cumbersome.Thus,inthismanuscriptwewillgiveanother
expressionofTSOTbyuseofdistributionfunctioninthespinor
Boltzmannequation(SBE).
TheSBEwasfirstlyproposedbyShengetalin1996atsteady
state[22],theyderivedthisequationfromKadanoffnonequilibrium
Greenfunction(NEGF)formalismbasedonthegradient
approximation[23].ThenSBEwasaccomplishedbyLevyetalin
2004,whichcouldillustratethetimedependentprocessofspintransporteffectively[24].SBEwasalsoextendedtothecasebeyond
gradientapproximationin2013[25].In2019,Wangetal.
successfullyincludedtheRashbaandDresslhausspin-orbitcoupling
intotheSBE[26],whichishelpfulforustoinvestigatetheSOT
fromthespindiffusionequation–thecontinuityequationforthe
spinaccumulationandspincurrent.Inthismanuscript,ourpurpose
istoinvestigateTSOTina2-dimensionalferromagnetbymeansof
SBE.WewillfindthatanunusualSOTdrivenbythetemperature
gradientinthesystemwithDresselhausspinorbitcoupling,itisjust
theTSOTwewant.
II.TheoreticalFormalism
Considerthespin-polarizedelectrontransportinatwo-
dimensionalferromagnetwithDresselhausSOC.It’sHamiltonian
canbewrittenas =−ℏ2
2 2 0+ ⋅ + ,whereis
electroniceffectivemass,Jisthes-dexchangecouplingconstant,
istheunitvectorforthelocalmagneticmomentofferromagnet,
0istheunitmatrix, isthePaulimatrixvector, describes
theDresselhausSOCintwodimensionalferromagnet.In1955,
DresselhausgavetheHamiltonianofDresselhausSOCinthree
dimensionalsystemwithabulkinversionasymmetry(BIA)[29].
Whensuchathreedimensionalsystemissubjectedtostrainatthe
interfaceorathinlayer,theHamiltonianwillreducetothefollowingsimplerform: =( − )[26],whereisthe
couplingconstantofDresselhausSOC,andarethexandy
componentsofgradientoperator,andthe and arethexand
ycomponentsofthePaulioperator .ByKadanoffnonequilibrium
Greenfunction(NEGF)formalism,wecanobtaintheSBEforthe
spinordistributionoftransportelectron[24]:
+ ⋅
− ⋅
+
ℏ , =−
d#1
where isthespinordistributionfunctionoftransportelectron,
whichcanbeexpandedtoa2×2matrix =↑↑↑↓
↓↑↓↓, isthe
spinorenergy,it’sdefinedas ()=() 0+1
2() (,).
IntheSBEframework,wecandecomposethespinordistribution
functionintotwopartsbyusingthecompletebasisofmatrices 0
andthe , ( , )=( , ) 0+ ( , ) ,where , is
scalardistributionfunctionand , isvectordistribution
function.Underthelocalequilibriumassumption,thespinor
distributionfunctioncanalsobewrittenasfollow:
, ,= 0 , +1 , 0+1 , ⋅ 2
where 0( , )isthelocalequilibriumdistributionfunction.
1( , )and1( , )arethenonequilibriumpartsofspinor
distribution.Intheferromagnet,wetakethelocalequilibrium
distributionfunctionasadiagonalmatrix 0( , )=↑↑0
0↓↓,herethediagonalcomponentsaretakenasFermidistribution
functions↑↑(↓↓)=()±1
2−()
()+1−1,whereis
Boltzmannconstant,()isthechemicalpotentialand≈
(FermiEnergy)atfinitetemperature.Wecanexpandthelocal
equilibriumdistributionbasedonthecompletebasis
0, , , ,whichis 0( , )=1
2(↑↑+↓↓) 0+1
2(↑↑−
↓↓) ,sothescalardistributionandvectordistributioncanbe
rewrittenas=1
2(↑↑+↓↓)+1and=1,1,1+1
2(↑↑−
↓↓).
Inspintronics,theSBEwithDresselhausspin-orbitcouplingina
two-dimensionalmagneto-electricsystemunderanexternalelectric
field hadbeengivenbyChaoYangetal.in2019[26]
+
⋅ − ⋅ , +
ℏ
−
=−
#3
and
+
⋅ − ⋅ , −
ℏ × , +
ℏ
−
+
ℏ2 − × , =−
#4
where−(
)and−(
)representthecollisionterms.
Undertherelaxationtimeapproximationassumption,wecan
derivetheequationsforthescalardistributionfunctionandthe
vectordistributionfunctionbasedonEqs.(3)and(4),whichcontainthelocalequilibriumdistributionfunctionas:
+
⋅ − ⋅ 1
2↑↑+↓↓+
+
⋅ − ⋅ 1
+
ℏ1
−1
=−−
#5
+
⋅ − ⋅ − × +
ℏ 1
2↑↑+↓↓
− 1
2↑↑+↓↓
+
ℏ 1
− 1
ℏ−
ℏ2 − × =− −
#6
whereandaretherelaxationtimesofelectronandspinflip,
respectively.Eq.(5)andEq.(6)arecoupledtogether,weshould
solvethemsimultaneously.Thephysicalobservablesinthe
spin-polarizedtransportcanbeexpressedbythesolutionsofscalar
andvectordistributionfunctions.Thechargedensityandcharge
currentaredefinedasfollow,
,=1 , , #7
and
,= 1 , , #8
whicharethemomentumintegralsoverthescalardistributions.
Similarly,thespinaccumulationandspincurrentdensityaredefine
as ,= , , #9
and
,= , , #10
whicharethemomentumintegralsoverthevectordistributions.It
shouldbenotedthatthespincurrent
( ,)isatensor.Moreover,
wecanalsodefinethethermalcurrentdensityas
,= 1 , , #11
whereisthescalarenergyofelectron.Sowhenwegetthe
solutionsofscalarandvectordistributionfunctions,wecanobtain
theseabovephysicalobservablesaccordingly.
Ontheotherhand,wecanalsoobtainthecontinuityequations
satisfiedbythesephysicalobservables.Ifweintegratethe
momentumovertheFermisurfaceonthebothsidesofEq.(5)and
Eq.(6),wehave
+ ⋅ =−
+
⋅ − ⋅ 1
2↑↑+↓↓
−
ℏ
−
−−
#12
and
+ ⋅ =
ℏ ×
−
ℏ 1
2↑↑+↓↓
− 1
2↑↑+↓↓
−
ℏ
−
+
ℏ2 − × =− −
#13
Eq.(12)isjustthecontinuityequationforchargedensityandcharge
current,whileEq.(13)isthecontinuityequationforthespin
accumulationandspincurrent,thelatterisalsocalledspindiffusion
equation.Whenthetime≫,theequationwillarriveatasteady
state,thenthespindiffusionequationwillreduceto
ℏ × = ⋅
+
ℏ 1
2↑↑+↓↓
− 1
2↑↑+↓↓
+
ℏ
−
−
ℏ2 − × #14
Fromtheabovesteadystateequation,wecanreadoutallthetorques
existinginthisspin-polarizedtransportprocess.Ontheleftsideof
thisequation,theterm
ℏ × isjustthespintransfertorque
givenbyLevyetal.[27].Ontherighthandsideside, ⋅
isthe
divergenceofspincurrent,whichalsocanmakeacontributiontothe
usualSTTasshownbyZhangetal[28].Besides,theterm
ℏ(
−
)−
ℏ2( − )× correspondstothe
usualspin-orbittorquepresentedbyWangetal.[26],whilethe
temperaturedependentterm
ℏ 1
2↑↑+↓↓
− 1
2↑↑+↓↓
isanewterm,itisinducedbythegradientof
localequilibriumdistributionfunction,werefertothisasthe
thermalSOT.Whenthegradientoftemperatureisappliedonly
alongx-direction,itcanbeexpressedas:
=−
ℏ.
1
2(−+1
2
)
1+(−−1
2
)2⋅−+1
2
+
(−−1
2
)
1+(−−1
2
)2⋅−−1
2
1
2
(15)
wecanseethatitisproportionaltothegradientoftemperature,
whichisconcededtothedefinitionofTSOTgivenbyFreimuthetal
[20-21],sothistermisjusttheTSOTwesearchfor,itisthecentral
resultinthismanuscript.Inthenext,wewillevaluatethesetorques
numericallyinaferromagnetwithDresselhausSOC.
III.NumericalResults
Weconsideratwo-dimensionalferromagnetwithDresselhaus
SOC,wherethesystemischosenasarectangularferromagnetwith
ageometryof25×25nm².Thetemperaturedistributionissimply
chosenas()=0+,whichislinearlydependentonthe
positionofxcomponent,where0isaconstant,kisthetemperaturegradient.FromEq.(15),wecanseethatthetemperature
gradientwillinducethermalspin-orbittorque.
Inordertoquantifythesetorquesandcurrents,weneedtosolve
Eq.(3)combiningwith(4)simultaneously,becausethescalar
distributionfunctionandvectordistributionfunctionarecoupled
togetherintheseequations.Tosimplifycalculation,wechosethe
unitvectorofmagnetizationasafixedvector =(0,0,1),the
equilibriumscalardistributionfunctionischosenas=1
2↑↑+
↓↓,andtheequilibriumvectordistributionfunctionisadoptedas
=((
+
),(
+
),(
+
))[22].Thedifferentialequations(3)and(4)aresolvedby
differencemethod.Thephysicalconstantandparametersarelisted
inTableⅠ,whereweadoptthematerialsparametersofferromagnet.
TableⅠ.Thephysicalconstantsandparameters
Physicalconstants/parametersSymbolValueUnit
Momentumrelaxationtime 10−13s
Spin-fliprelaxationtime 10−12s
Fermienergy 4 eV
Fermiwavevector 1.02×1010−1
s-dexchangecouplingstrengthJ 0.1 eV
Electricalfield E −5×104.−1Temperaturegradient 5×109.−1
InFig.1,weplotthechargecurrentdensityasafunctionof
positionxandy.Thevariationofchargecurrentwithrespectto
positionandtimeisgovernedbythecontinuityequationofcharge
densityandchargecurrentdensity(12).Forsimplicity,weonly
studythechargecurrentatsteadystate.Wecanseethatthecharge
currentdensitydecreasesgraduallyalongboththexandydirections,
whichisduetotheresistanceintheferromagnet,inourcalculationit
isconcernedwiththemomentumrelaxationtimeofelectrons.
Sincetheexternalelectricfieldisappliedonlyalongthex-direction,
thevariationofchargecurrentalongy-directionismainlycausedby
theDresselhausSOC.
Fig.1ThechargecurrentdensityFig.2Thethermalcurrentdensity
vsposition vsposition
Wealsoshowthecurveofthermalcurrentdensityasafunction
ofpositioninFig.2,whichissimilartothechargecurrentdensity
becauseoftheirdefinitions,italsodecreaseswithpositiongradually.
Hereweonlyconsiderthethermalcurrentdensitycarriedbythetransportelectrons.Besidestheexternalelectricfield,thethermal
currentdensitycanalsobedrivenbythetemperaturegradient.Since
theelectricfieldandgradientoftemperatureareallalongthex-axis,
thevariationofthermalcurrentdensityalongy-directionisprimarily
inducedbytheDresselhausSOC.
Becausewechoosethemagnetizationofferromagnet =
(0,0,1),sothezcomponentofSTTis0.InFig.3,weshowthexand
ycomponentsofSTTdensityasafunctionofposition.Theusual
STTisthespaceintegralofthisdensityovertherectangular
ferromagnet.ItisshownthatthemagnitudeofSTTdensityis
differentatdifferentpositionforboththexandycomponents.The
magnetizationofferromagnetat=10iseasiesttobe
switchedbythebiggerSTT,andishardesttofliparoundtheline
=becauseofthesmallerSTT.Theswitchingofmagnetization
willproducethespinwavewithinthetwo-dimensionalferromagnet.
Fig.3(a)Thex-componentofSTTdensity.(b)They-componentofSTTdensity
ThespincurrentdensityvspositionisshowninFig.4.Sincethespincurrentisatensor,weonlydrawthexx-,xy-andxz-
componentsofspincurrentdensityasafunctionofposition,they
varyobviouslyaroundtheline=.Thevariationofspincurrent
withpositionandtimesatisfiesthecontinuityequation(13)forthe
spinaccumulationandspincurrent.AccordingtoEq.(14),the
divergenceofspincurrentwillmakeacontributiontotheZhang-like
STT[28].
Fig.4(a)Thexx-componentofspincurrent(b)Thexy-componentofspin
current(c)Thexz-componentofspincurrent
InFig.5,wedrawtheSOTasafunctionofposition.Theusual
SOTisexpressedas
ℏ(
−
)−
ℏ2( − )× ,
soitisnotsensitivetothetemperature.Itdecreasesalongx-
directionobviously,whilethereissmallvariationalongy-direction,
becausetheexternalelectricfieldisappliedalongx-direction.For
comparison,wealsoplottheTSOTatdifferenttemperature300K,
200Kand100KinFig.6,respectively,whichdependonthe
temperatureandit’sgradientobviously.Thehigheroftemperature,
thebiggerofTSOT,becausetherearemorepolarizedelectronFig.5TheSOTvsposition.Fig.6TheTSOTvspositionatdifferent
temperatureT=300K,200K,100K.
participatingintransportathighertemperature.ComparedFig.6
withFig.5,wecanfindthattheTSOTissmallerthantheusualSOT,
whiletheTSOTcanbecomebiggerwhenweincreasethe
temperature,sotheTSOTcannotbenegligibleathigher
temperature.Certainly,TSOTisalsoproportionaltothetemperature
gradient,itwillplayanimportantroleathighertemperaturegradient.
ItshouldbepointedoutthattheTSOTiscalculatedbyEq.(15),we
onlyneedtheexpressionoflocalequilibriumdistributionfunction,it
isverysimplethanFreimuth’sexpressionofBerryphase[20-21],
becausethelatterneedtheelectronicwavefunctionobtainedusually
byDFT.BymeansofEq.(15),wecancalculatetheTSOTeasier,
thisistheadvantageofSBEmethod.
IV.SummaryandDiscussions
Inthispaper,wehavederivedtheTSOTinatwo-dimensional
ferromagnetwithDresselhausSOCbySBEunderthelocal
equilibriumassumption.TheusualSOTisinducedbytheexternal
electricfieldappliedtothesystem,whiletheTSOTisdrivenbythegradientoftemperature.WealsofindthatTSOTisverysensitiveto
thetemperature,thehighertemperature,thebiggerTSOT.Our
resultsshowthattheTSOTissmallerthanSOT,butitcannotbe
negligibleathighertemperature.Certainly,TSOTisalso
proportionaltothetemperaturegradient,accordingtoitsexpression
Eq.(15).BecausethedirectexperimenttoobserveTSOTin
two-dimensionalferromagnetswithDresselhausspin-orbitcoupling
haven’tbeencarriedoutnow,weonlypredicttheoreticallythatone
canobservetheeffectsofTSOTinthecaseofbiggradientof
temperatureandhighertemperature.
Tosimplifyourcalculation,weonlychooseasimpleuniform
magnetizationina2-dimensionalferromagnet,whileinrealitythe
magnetizationusuallyvarieswithtimeandposition.Thevariationof
magnetizationwouldhaveinfluenceonthetransportpropertiesof
thespin-polarizedelectrons.Ifweconsiderthevariationof
magnetization,thecalculationwillbecomemuchmorecomplicated,
itisleftforfutureexploration.
Acknowledgments
ThisstudyissupportedbytheNationalKeyR&DProgramof
China(GrantNo.2022YFA1402703),theStrategicPriority
ResearchProgramoftheChineseAcademyofSciences(GrantNo.
XDB28000000).WealsothankProf.GangSu,Zhen-Gang.Zhu,Bo.GuandQing-Bo.Yanfortheirhelpfuldiscussions.
AUTHORCONTRIBUTIONSTATEMENT
Inthiswork,Zheng-ChuanWangproposedtheidea,Chun-Yi
Xueperformedthecalculation,analyzedthenumericalresults,and
wrotethemanuscript.Ya-RuWangassistedwiththecalculationand
analysis.
DataAvailabilityStatement
Datasetsgeneratedduringthecurrentstudyareavailablefromthe
correspondingauthoronreasonablerequest.
References
[1]A.ManchonandS.Zhang,Phys.Rev.B78,212405(2008).
[2]T.KuschelandG.Reiss,Nat.Nanotechnol.10,22(2015).
[3]A.Matos-AbiagueandR.L.Rodrıguez-Suarez,Phys.Rev.B80,
094424(2009).
[4]K.M.D.HalsandA.Brataas,Phys.Rev.B88,085423(2013).
[5]A.Brataas,A.D.KentandH.Ohno,Nat.Mater.11:372-81
(2012).
[6]M.N.Baibichetal.,Phys.Rev.Lett.61,2472(1988).
[7]J.C.Slonczewski,J.Magn.Magn.Mater.159,L1(1996).
[8]W.J.GallagherandS.S.P.Parkin,IBMJ.Res.Dev.50,5
(2006).
[9]X.Hanetal.,Appl.Phys.Lett.118,120502(2021).[10]S.Aggarwaletal.,2019IEEEInt.ElectronDevicesMeet.
(IEDM)pp.2.1.1–2.1.4(2019).
[11]L.Savtchenkoetal.,USPatent6546906B1(2003).
[12]M.Motoyoshietal.,Dig.Tech.Pap.-Symp.VlSITechnol.p.22
(2004).
[13]C.Chappert,A.FertandF.N.Van-Dau,Nat.Mater.6:813–23
(2007).
[14]T.Kishietal.,Tech.Dig.-Int.ElectronDevicesMeet.309
(2008).
[15]R.Patel,E.Ipek,E.G.Friedman,MicroelectronicsJournal,
45(2),133-143(2014).
[16]C.Bi,N.SatoandS.X.Wang,AdvancesinNon-Volatile
MemoryandStorageTechnology,203-235(2019).
[17]J.M.Kimetal.,NanoLett.20(11):7803-7810(2020).
[18]D.J.Kimetal.,Nat.Commun.8,1400(2017).
[19]M.HatamiandG.E.W.Bauer,Phys.Rev.Lett.99,066603
(2007).
[20]F.Freimuth,S.BlugelandY.Mokrousov,J.Phys:Condens.
Matt.26,104202(2014).
[21]G.Geranton,F.Freimuth,S.BlugelandY.Mokrousov,Phys.
Rev.B91,014417(2015).
[22]L.Sheng,D.Y.Xing,Z.D.Wang,andJ.Dong,Phys.Rev.B55,5908(1997).
[23]L.P.Kadanoff,G.Baym,andJ.D.Trimmer,Quantum
StatisticalMechanics(W.A.Benjamin,1962).
[24]J.Zhang,P.M.Levy,S.ZhangandV.Antropov,Phys.Rev.Lett.
93,256602(2004).
[25]Z.C.Wang,TheEuro.Phys.Jour.B86,206(2013).
[26]C.Yang,Z.C.Wang,Q.R.ZhengandG.Su,TheEuro.Phys.
Jour.B92,136(2019).
[27]A.S.Shpiro,P.M.Levy,S.Zhang,Phys.Rev.B67,104430
(2003).
[28]S.ZhangandZ.Li,Phys.Rev.Lett.93,27204(2004).
[29]G.Dresselhaus,A.F.Kip,andC.Kittel,Phys.Rev.
100,618(1955). |
1203.4079v2.Spin_orbit_couplings_between_distant_electrons_trapped_individually_on_liquid_helium.pdf | arXiv:1203.4079v2 [quant-ph] 13 Nov 2012Spin-orbit couplings between distant electrons trapped in dividually on
liquid Helium
M. Zhang1and L. F. Wei∗1,2
1Quantum Optoelectronics Laboratory, School of Physics,
Southwest Jiaotong University, Chengdu 610031, China
2State Key Laboratory of Optoelectronic Materials and Techn ologies,
School of Physics and Engineering, Sun Yat-sen University, Guangzhou 510275, China
(Dated: November 10, 2018)
Abstract
We propose an approach to entangle spins of electrons floatin g on the liquid Helium by coherently ma-
nipulating their spin-orbit interactions. Theconfigurati on consists ofsingleelectrons, confinedindividually
on liquid Helium by the micro-electrodes, moving along the s urface as the harmonic oscillators. It has
been known that the spin of an electron could be coupled to its orbit (i.e., the vibrational motion) by prop-
erly applying a magnetic field. Based on this single electron spin-orbit coupling, here we show that a
Jaynes-Cummings (JC) type interaction between the spin of a n electron and the orbit of another electron at
a distance could be realized via the strong Coulomb interact ion between the electrons. Consequently, the
proposed JC interaction could be utilized to realize a stron g orbit-mediated spin-spin coupling and imple-
ment the desirable quantum information processing between the distant electrons trapped individually on
liquid Helium.
PACSnumbers: 73.20.-r, 03.67.Lx, 33.35.+r
∗weilianfu@gmail.com
1I. INTRODUCTION
The interactions between the microscopic particles, e.g., the ions in Paul trap [1], the neutral
atoms confined in optical lattice [2], and the electrons in Pe nning trap [3], etc., relate usually to
their masses and the inter-particle forces. Due to the small mass and the strong Coulomb interac-
tion,theinteractingelectronscouldbeused toimplementq uantuminformationprocessing(QIP).
Theideaofquantumcomputingwithstrongly-interactingel ectronsonliquidHeliumwasfirstpro-
posedbyPlatzmanandDykmanin1999[4]. Intheirproposal,t hetwolowerhydrogen-likelevels
of the surface-state electron are encoded as a qubit, and the effectively interbit couplings can be
realizedbytheelectricdipole-dipoleinteraction. Whent heliquidheliumiscooledontheorderof
mK temperature the qubit possesses long coherent time (e.g. , up to the order of ms) [5, 6]. Inter-
estingly,Lyonsuggested[7]thatthequbitscouldalsobeen coded by thespinsoftheelectrons on
liquidHelium,andestimatedthatthequbitcoherenttimeco uldreach 100s[7]. Heshowedfurther
that the magneticdipole-dipoleinteractions between the s pinscould be used to couplethe qubits,
if the electrons are confined closed enough. For example, the coupling strength can reach to the
order of kHz for the distance d= 0.1µm between the electrons [7]. Remarkably, recent experi-
ments[8–10]demonstratedthemanipulationsofelectrons( confining,transporting,anddetecting)
onliquidHeliuminthesingle-electronregime. Thisprovid esreallytheexperimentalplatformsto
realizetherelevantQIPwithelectrons onliquidHelium[11 –14].
Here, we propose an alternative approach to implement QIP wi th electronic spins on liquid
Heliumbycoherentlymanipulatingthespin-orbitinteract ionsoftheelectrons. Inourproposal,the
virtuesoflong-livedspinstates(toencodethequbit)ands trongCoulombinteraction(forrealizing
theexpectably-fastinterbitoperations)arebothutilize d. Theelectronsaretrappedindividuallyon
the surface of liquid Helium by the micro-electrodes. In the plane of liquid Helium surface each
electron moves as a harmonic oscillator. It has been showed t hat such an external orbit-vibration
could be effectively coupled to the internal spin of a single electron by applying a magnetic field
with a gradient along the vibrational axis [13]. Interestin gly, we show that the spin of an electron
could be coupled to the vibrational motion of another distan t electron [as a Jaynes-Cummings
(JC) type interaction], by designing a proper virtual excit ation of the electronic vibration. The
present JC interaction could be utilized to significantly en hance the spin-spin coupling between
the distant electrons, and implement the desirable quantum computation with the spin qubits on
liquidHelium.
2Qe
HLiquid Helium xz
y
+I
BzBsPotential h
FIG. 1: (Color online) Sketch for a single electron trapped o n the surface of liquid Helium. The liquid
Helium provides z-directional confinement, and the micro-electrode Q (below the Helium surface at depth
H) traps the electron in x-yplane. The desirable spin qubit is generated by an applied un iform magnetic
fieldBs, and the spin-orbit coupling of the trapped electron is obta ined by applying a current to another
micro-electrode I (upon the liquid Helium surface at the hei ghth).
Thepaperisorganizedasfollows: InSec. IIwediscusstheme chanismforspin-orbitcoupling
withasingleelectrontrappedonliquidHelium[13],andthe nshowhowtoutilizesuchacoupling
to realize the desirable quantum gate with the single electr on. By using the electron-electron
Coulomb interaction, in Sec. III, we propose an approach to i mplement the JC coupling between
the spin of an electron and the orbital motion of another elec tron. Based on such a distant spin-
orbit interaction, we show that a two-qubit controlled-NOT (CNOT) gate and an orbit-enhanced
coupling between the distant spins could be implemented. Fi nally, we give a conclusion in Sec.
IV.
II. SPIN-ORBITCOUPLINGWITHA SINGLETRAPPEDELECTRON
We consider first a single electron trap shown in Fig. 1 [13], w herein an electron (with mass
meand charge e) on liquid Helium is weakly attracted by its dielectric imag e potential V(z) =
−Λe2/z(withΛ = (ε−1)/4(ε+ 1)andεbeing the dielectric constant of liquid Helium). Due
tothePauliexclusionprinciple,thereisanbarrier(about 1eV)topreventtheelectronpenetrating
into the liquid Helium. As a consequence, z-directional confinement of the electron is realized,
yielding an one-dimensional (1D) hydrogenlike atom with th e spectrum En=−/planckover2pi1R/n2[15].
Here,R= Λ2e4me/(2/planckover2pi12)≈170GHz and rb=/planckover2pi12/(mee2Λ)≈7.6nm are theeffectiveRydberg
3energy and Bohr radius, respectively. In x-yplane, the electron can be confined by the micro-
electrode Q located at Hbeneath the liquid Helium surface. Typically, x,y,z≪H, and thus the
potentialoftheelectron can bedescribed by[5]
U(x,y,z)≈ −Λe2
z+E⊥z+me
2(ν2
xx2+ν2
yy2) (1)
withE⊥=eQ/H2,νx=νy=/radicalbig
eQ/(meH3), andQbeing the effective charge of the micro-
electrode. This potential indicates that the motions of the trapped electron are a 1D Stark-shifted
hydrogen along the z-direction, and a 2D harmonic oscillator in the plane parall el to the liquid
Heliumsurface. TheHamiltonianfortheorbitalmotionsoft hetrapped electron can bewrittenas
ˆHo=/summationdisplay
nEn|na/an}bracketri}ht/an}bracketle{tna|+/summationdisplay
k=x,y/planckover2pi1νk(ˆa†
kˆak+1
2). (2)
Here,|na/an}bracketri}htis thenth boundstateofthehydrogenlikeatom, ˆa†
kandˆakare thebosonicoperatorsof
thevibrationalquantaoftheelectron alongthe k-direction.
A spin qubit is generated by applying an uniform magnetic fiel dBsalongxdirection, and
its Hamiltonian reads ˆHq= (gµBBs)ˆσx/2. Here, the Pauli operator is defined as ˆσx=| ↑/an}bracketri}ht/an}bracketle{t↑
| − | ↓/an}bracketri}ht/an}bracketle{t↓ | with| ↓/an}bracketri}htand| ↑/an}bracketri}htbeing the two spin states. g= 2is the electronic g-factor, and
µB= 9.3×10−24J/T is the Bohr magneton. The spin-orbit coupling of the trap ped electron can
be realized by applying a dc current Ito the electrode I (located upon the liquid Helium surface
with a height h) [13]. Typically, x,z≪hand the magnetic field generated by the current I
reads/vectorB= (Bx,0,Bz)withBx≈µ0I(1−z/h)/(2πh)andBz≈µ0Ix/(2πh2). Here,µ0is
the permeability of free space. Therefore, the Hamiltonian describing theinteraction between the
magnetic field and spin can be expressed as: ˆHsb=gµB(Bzˆσz+B′
xˆσx)/2withB′
x=Bs+Bx,
ˆσz= ˆσ−+ ˆσ+,ˆσ−=| ↓/an}bracketri}ht/an}bracketle{t↑ |andˆσ+=| ↑/an}bracketri}ht/an}bracketle{t↓ |. Consequently, the total Hamiltonian of the
trapped electronin theappliedmagneticfields reads
ˆH=/planckover2pi1νs
2ˆσx+ˆHo+ˆHsx, (3)
with
ˆHsx=gµBµ0I
4πh2/radicalbigg
/planckover2pi1
2meνx(ˆax+ˆa†
x)ˆσz. (4)
ThefirstandsecondtermsintherighthandofEq.(3)describe thefreeHamiltonianofthetrapped
electron,with νs= (gµB//planckover2pi1)[Bs+(µ0I/2πh)]beingthetransitionfrequencybetweenitstwospin
states, and ˆHsxdescribes the coupling between the spin and the orbital moti on along x-direction.
4Note that the coupling between the spin and z-directional orbital motion is neglected, due to the
large-detuning. Also, theappliedstrongfield Bs(e.g.,0.06T)does notaffect theinteraction ˆHsx,
althoughitwillchangeslightlytheelectron’s motionsint hey-zplane[16].
Obviously,theHamiltonianin Eq.(3)can besimplifiedas
ˆHe=/planckover2pi1Ω/parenleftbig
eiδtˆσ+ˆa+e−iδtˆσ−ˆa†/parenrightbig
(5)
intheinteractionpicture. Here, δ=νs−νxisthedetuning,
Ω =gµBµ0I
4πh2√2/planckover2pi1meνx(6)
is the coupling strength, and ˆa= ˆax,ˆa†= ˆa†
x. Note that, the Hamiltonian in Eq. (5) can also
be obtained by applying an ac current I(t) =Icos(ωt)with frequency ω=νx−νs+δto the
electrode. Specially, when δ= 0, this Hamiltonian describes a JC-type interaction between the
spin and orbit of the single electron. In fact, Ref. [13] has a rranged this spin-orbit coupling of
a single electron to increase the interaction between the sp in and a quantized microwave field.
Alternatively, we will utilize this spin-orbit coupling (t ogether with the electron-electron strong
Coulomb interaction) to realize a strong interaction betwe en two electronic spins and generate
certain typicalquantumgates.
For the typical parameters: I= 1mA,h= 0.5µm, andνx= 10GHz [5, 13], we have
Ω≈5.2MHz. Thisissignificantlylargerthanthedecoherencerate( whichistypicallyontheorder
of10kHz [5, 13]) of thevibrational states of thetrapped electro n. Thus, the aboveJC interaction
providesapossibleapproachtoimplementQIPbetweenthesp inandorbitstatesofasingletrapped
electron. FortheJCinteraction,thestate-evolutionscan belimitedintheinvariant-subspaces {| ↓
,0/an}bracketri}ht}and{| ↓,1/an}bracketri}ht,| ↑,0/an}bracketri}ht},with|0/an}bracketri}htand|1/an}bracketri}htbeingthegroundandfirstexcitedstatesoftheharmonic
oscillator. Thus,aphasegate ˆP=|0,↓/an}bracketri}ht/an}bracketle{t0,↓ |+|0,↑/an}bracketri}ht/an}bracketle{t0,↑ |+|1,↓/an}bracketri}ht/an}bracketle{t1,↓ |−|1,↑/an}bracketri}ht/an}bracketle{t1,↑ |couldbe
implementedbyapplyingacurrentpulsetotheelectrodeI. T herelevantduration tissettosatisfy
theconditions: sin(Ωt)≈0andcos(√
2Ωt)≈ −1(e.g.,Ωt≈37.7numerically). Consequently,a
CNOTgatewiththesingleelectroncouldberealizedas ˆS=ˆR(π/2,−π/2)ˆPˆR(π/2,π/2),where
ˆR(α,β) = (| ↑/an}bracketri}ht/an}bracketle{t↑ |+| ↓/an}bracketri}ht/an}bracketle{t↓ |)cos(α)−i[exp(iβ)| ↑/an}bracketri}ht/an}bracketle{t↓ |+exp(−iβ)| ↓/an}bracketri}ht/an}bracketle{t↑ |]sin(α)isanarbitrary
single-bit rotation [17]. This CNOT gate operation, betwee n the spin states and the two selected
vibrational states of a single electron [18], is an intermed iate step for the later CNOT operation
between twodistantspinqubits.
5Q1 Q2e1 e2
d
Liquid Helium
+I1
+I2
H Hh h
Potential Bsz
x y
FIG. 2: (Color online) Two electrons (denoted by e1ande2) are confined individually in two potential
wells with the distance d, which is sufficiently large (e.g., d= 10µm) such that the magnetic dipole-
dipole coupling between theelectronic spins isnegligible . Theorbital motionsof thetwoelectrons arealso
decoupled from each other, since they are trapped in large-d etuning regime. By applying a current to the
electrode I 1the spin of the electron e1could be coupled to the vibrational motions of electron e2, via a
virtual excitation of the vibrational motion of electron e1.
III. SPIN-ORBITJC COUPLINGBETWEENTHEDISTANTELECTRONS
Without loss of generality, we consider here two electrons ( denoted by e1ande2) trapped
individually in two potential wells, see Fig. 2. Suppose tha t the distance dbetween the potential
wells is sufficiently large (e.g., d= 10µm), such that the directly magnetic interaction between
the two spins could be neglected. Thus, the interaction betw een the two electrons leaves only
the Coulomb one. Specially, the Coulomb interaction along t hex-direction can be approximately
writtenas
V(x)≈e2
2πǫ0d3x1x2 (7)
withxjbeing the displacement of electron ejfrom its potential minima. By controlling the volt-
ages applied on the electrodes Q1andQ2, the vibrational frequencies of the electrons are set as
thelarge-detuning(and thustheelectrons aredecoupled fr omeach other).
To couple the initially-decoupled electrons, we apply a cur rentIto the electrode I 1. As dis-
cussed previously,such acurrent induces aspin-orbitcoup ling[i.e., ˆHein Eq. (5)]of theelectron
e1. Therefore, thepresenttwo-electronssystemcan bedescri bed bythefollowingHamiltonian
ˆHee=ˆHe+/planckover2pi1˜Ω/parenleftBig
ei∆tˆaˆb†+e−i∆tˆa†ˆb/parenrightBig
(8)
in the interaction picture. Where, ˆbandˆb†are the bosonic operators of the vibrational motion of
electrone2alongx-direction, ∆ =ν2x−ν1xisthedetuningbetweenthetwoelectronicvibrations
6alongx-direction, and
˜Ω =e2
4πǫ0med3√ν1xν2x, (9)
the coupling strength. Numerically, for d= 10µm andνjx= 10GHz we have ˜Ω≈25MHz.
Above,thespinofelectron e2wasdropped,asthedriving(inducedbyelectrode I1)onthisspinis
negligible(dueto d≫h).
The dynamical evolution ruled by the Hamiltonian in Eq. (8) i s given by the following time-
evolutionoperator
ˆU(t) = 1+/parenleftbig−i
/planckover2pi1/parenrightbig/integraltextt
0ˆHee(t1)dt1
+/parenleftbig−i
/planckover2pi1/parenrightbig2/integraltextt
0ˆHee(t1)/integraltextt1
0ˆHee(t2)dt2dt1+···.(10)
Weassume δ= ∆forsimplicity,thentheabovetime-evolutionoperatorcan beapproximatedas
ˆU(t)≈exp/parenleftbigg
−it
/planckover2pi1ˆHeff/parenrightbigg
, (11)
withtheeffectiveHamiltonian
ˆHeff=/planckover2pi1Ω2
δ/bracketleftbig
ˆa†ˆa(ˆσ+ˆσ−−ˆσ−ˆσ+)+ ˆσ+ˆσ−/bracketrightbig
+/planckover2pi1˜Ω2
δ/parenleftBig
ˆb†ˆb−ˆa†ˆa/parenrightBig
+/planckover2pi1Ω˜Ω
δ/parenleftBig
ˆσ+ˆb+ ˆσ−ˆb†/parenrightBig
.(12)
Thesecondterm intherighthand ofEq.(10)andthetermsrela tingto thehighorders of Ω/δand
˜Ω/δwere neglected,since Ω,˜Ω≪δ. Furthermore, at theexperimentaltemperature(e.g., 20mK)
the electrons are frozen well into their vibrational ground states (about 40mK for the vibrational
frequency ∼10GHz). Thismeansthattheexcitationofthevibrationofelec trone1isvirtual,and
thus the terms in Eq. (12) related to ˆa†ˆacan be adiabatically eliminated. As a consequence, the
HamiltonianinEq. (12)reduces to
ˆHeff=/planckover2pi1Ω2
δˆσ+ˆσ−+/planckover2pi1˜Ω2
δˆb†ˆb+/planckover2pi1Ω˜Ω
δ/parenleftBig
ˆσ+ˆb+ ˆσ−ˆb†/parenrightBig
(13)
and furtherreads (for Ω =˜Ω)
ˆHJC=/planckover2pi1Ω2
δ/parenleftBig
ˆσ+ˆb+ ˆσ−ˆb†/parenrightBig
(14)
in the interaction picture. Obviously, this Hamiltonian de scribes a JC-type coupling between the
spinofelectron e1and theorbitalmotionofelectron e2.
70 2 4 6 810 1200.20.40.60.81
t (s)Occupancies |↑1,01,02〉 |↓1,01,12〉
×10−7
FIG. 3: (Color online) Numerical solutions for the Hamilton ian in Eq. (8): the occupancy evolutions of
states| ↑1,01,02/an}bracketri}ht(bluecurve)and | ↓1,01,12/an}bracketri}ht(redcurve),with ˜Ω = Ω = 25 MHzand δ= ∆ = 250 MHz.
Typically, the effective coupling strength can reach Ω′= Ω2/δ≈2.5MHz for d= 10µm,
ν1x= 10GHz, and δ= 250MHz. With these parameters and the Hamiltonian in E.q (8), Fi g. 3
showsnumerically theoccupancy evolutionsof thestates | ↑1,01,02/an}bracketri}htand| ↓1,01,12/an}bracketri}ht. Here,| ↓j/an}bracketri}ht
and| ↑j/an}bracketri}htare the two spin states of electron ej, and|0j/an}bracketri}htand|1j/an}bracketri}htare the two lower vibrational
states of the electron. Obviously, the results are well agre ement with the solutions (i.e., the time-
dependentoccupanciesof | ↑1,02/an}bracketri}htand| ↓1,12/an}bracketri}ht)fromtheHamiltonian ˆHJC. Thisverifiesthevalid-
ityofˆHJC. Thespin-orbitJCcoupling(14)couldbeusedtoimplementQ IPbetweentheseparately
trapped electrons. For example, by applying a current pulse with the duration t=π/(2Ω′)to an
electrode, e.g., I 1, a two-qubit operation ˆV1,2(π/2) =| ↓1,02/an}bracketri}ht/an}bracketle{t↓1,02|−i| ↓1,12/an}bracketri}ht/an}bracketle{t↑1,02|between
the electrons could be implemented. Consequently, a CNOT ga te between the qubits encoded by
theelectronicspinscouldbeimplementedbytheoperationa lsequence ˆC=ˆV1,2(π/2)ˆS2ˆV1,2(π/2),
withˆS2being the single-electron CNOT gate operated on the electro ne2. After this two-spin
CNOT operation, the vibrational motions of the trapped elec trons return to their initial ground
states.
Furthermore, the mechanism used above for the distant spin- orbit coupling can be utilized
to implement an orbit-mediated spin-spin interaction, whe rein the degrees freedom of the orbits
of the two electrons are adiabatically eliminated. Indeed, by applying the current pulses to the
electrodes simultaneously,theHamiltonianoftheindivid ually-drivenelectrons reads:
ˆH′
ee=/planckover2pi1Ω/parenleftbig
eiδtˆσ+ˆa+e−iδtˆσ−ˆa†/parenrightbig
+/planckover2pi1˜Ω/parenleftBig
eiδtˆaˆb†+e−iδtˆa†ˆb/parenrightBig
+/planckover2pi1G/parenleftBig
eiηtˆτ+ˆb+e−iηtˆτ−ˆb†/parenrightBig
.(15)
80 20 40 60 80 100 12000.20.40.60.81
t (s)Occupancies
× 10−6|↓1,01,02,↑2〉 |↑1,01,02,↓2〉
FIG. 4: (Color online) Numerical solutions for the Hamilton ian in Eq. (15): the occupancy evolutions of
the states | ↓1,01,02,↑2/an}bracketri}ht(blue curve) and | ↑1,01,02,↓2/an}bracketri}ht(red curve), with ˜Ω = 25MHz,Ω = 2.6MHz,
δ= 250MHz,and η= Ω2/δ.
Here, the first and third terms describe respectively the spi n-orbit couplings of the electrons e1
ande2, and the second term describes the Coulomb interaction betw een the electrons. Gandη
are the coupling strength and the detuning between the spin a nd orbital motions of electron e2,
respectively. ˆτ−=| ↓2/an}bracketri}ht/an}bracketle{t↑2|andˆτ+=| ↑2/an}bracketri}ht/an}bracketle{t↓2|are the corresponding spin operators of electron
e2. The spin-orbit couplings, i.e., the first and third terms in the Hamiltonian, can be realized by
applying the ac currents I1(t) =I1cos(ω1t)andI2(t) =I2cos(ω2t)to the electrodes I 1and I2
respectively,with thefrequencies ω1=ν1x−νs+δandω2=ν2x−νs+η. Here the ac currents
are appliedtorelatively-easilysatisfytheaboverequire ments forthedetunings.
Withthehelp ofEq.(13), Eq.(15)can beeffectivelysimplifi edas
ˆH′
ee=ˆHeff+/planckover2pi1G/parenleftBig
eiηtˆτ+ˆb+e−iηtˆτ−ˆb†/parenrightBig
, (16)
i.e.,
ˆH′
ee=/planckover2pi1Ω˜Ω
δ/parenleftBig
eiγtˆσ+ˆb+e−itγˆσ−ˆb†/parenrightBig
+/planckover2pi1G/parenleftBig
ei(η−˜Ω2/δ)tˆτ+ˆb+e−i(η−˜Ω2/δ)tˆτ−ˆb†/parenrightBig
(17)
in the interaction picture, with γ= (Ω2−˜Ω2)/δ. We select G= Ω˜Ω/δandη= Ω2/δfor
simplicity,suchthat
ˆH′
ee=/planckover2pi1G/parenleftBig
eiγtˆσ+ˆb+e−itγˆσ−ˆb†/parenrightBig
+/planckover2pi1G/parenleftBig
eiγtˆτ+ˆb+e−iγtˆτ−ˆb†/parenrightBig
. (18)
ByrepeatingthesamemethodforderivingtheeffectiveHami ltonianˆHeff,i.e.,neglectingtheterms
relatingto thehighorders of G/γinthetime-evolutionoperatorand eliminatingadiabatica llythe
9termsrelating to ˆb†ˆb, wehave
ˆH′
eff=/planckover2pi1G2
γ(ˆσ+ˆτ−+ ˆσ−ˆτ+). (19)
This is an effectively interaction between the two spins, me diated by their no-excited orbital mo-
tions[19].
Numerically, for ˜Ω≈25MHz,Ω≈2.6MHz, and δ≈250MHz, we have |γ| ≈2.5MHz,
G≈0.26MHz, and Ω′′=|G2/γ| ≈27kHz. With these parameters, Fig. 4 shows numerically
thetime-dependentoccupanciesof | ↓1,01,02,↑2/an}bracketri}htand| ↑1,01,02,↓2/an}bracketri}htfromtheHamiltonianinEq.
(15). This provides the validity of the simplified Hamiltoni an in Eq. (19). Obviously,the present
orbit-mediated spin-spin coupling is significantly weaker than the above spin-orbit JC coupling
(14) between the electrons, but still stronger than the dire ctly magnetic dipole-dipole coupling
(which is estimated as ∼10−3Hz for the same distance) between the spins. Since the cohere nce
time of the spin qubit is very long (e.g., could be up to minute s [7]), the orbit-mediated spin-
spin coupling demonstrated above could be utilized to gener ate the spins entanglement and thus
implementthedesirableQIP.
Finally, we would like to emphasize that, the considered dou ble-trap configuration shown in
Fig. 2 seems similarly to that of the recent ion-trap experim ents [20, 21]. There, two ions are
confinedintwopotentialwellsseparatedby 40µm[20](or 54µm[21]),andtheion-ionvibrational
coupling ˆHii=/planckover2pi1˜Ω[exp(i∆t)ˆaˆb†+ exp(−i∆t)ˆa†ˆb]is achieved up to ˜Ω≈10kHz [20] (or ˜Ω≈
7kHz [21]). The coupling between theions was manipulated tun ablyby controllingthe potential
wells(viasweepingthevoltagesontherelevantelectrodes )toadiabaticallytunetheoscillatorsinto
or out of resonance, i.e., ∆ = 0or∆≫˜Ω[22], respectively. Instead, in the present proposal we
suggested a JC-type coupling (and consequently an orbit-me diated spin-spin coupling) between
the two separated electrons. Therefore, the operational st eps for implementing the QIP should be
relativelysimple. Moreinterestingly,heretheelectron- electroncouplingstrength ˜Ωissignificantly
stronger (about 103times) than that between the trapped ions (e.g,9Be+[20]), since the mass of
electron ismuchsmallerthanthatoftheions.
IV. CONCLUSION
We have suggested an approach to implement the QIP with elect ronic spins on liquid helium.
Twolong-livedspinstatesofthetrappedelectronwereenco dedasaqubit,andthestrongCoulomb
10interactionbetweentheelectronswasutilizedasthedatab us. Thespin-orbitJCcouplingbetween
the spin of an electron and the vibrational motion of another distant electron is generated by
designing a virtual excitation of the electronic vibration . Such a distant spin-orbit interaction
is further utilized to realize an orbit-mediated spin-spin coupling and implement the desirable
quantumgates.
Compared with the ions in the Paul traps, here a feature is tha t the mass of the electron is
much smaller than that of ions, and thus a strong Coulomb coup ling up to 25 MHz between the
electrons could reached for a distance of d= 10µm. Finally, the construction suggested here
for implementing quantum computation with trapped electro ns on the liquid helium should be
scalable, andhopefullybefeasiblewithcurrent micro-sca letechnique.
Acknowledgements : This work was partly supported by the National Natural Scie nce
FoundationofChinaGrantsNo. 11204249,11147116,1117437 3,and90921010,theMajorState
Basic Research Development Program of China Grant No. 2010C B923104, and the open project
ofStateKeyLaboratoryofFunctionalMaterials forInforma tics.
[1] W.Paul,Rev.Mod.Phys. 62,531(1990); D.J.Wineland, C.Monroe, W.M.Itano, D.Leibfr ied, B.E.
King,andD.M.Meekhof, J.Res.Natl.Inst.Stand.Technol. 103, 259(1998); J.I.CiracandP.Zoller,
Phys. Rev. Lett. 74, 4091 (1995).
[2] O.Morsch and M.Oberthaler, Rev.Mod. Phys. 78, 179 (2006).
[3] L. S. Brownand G. Gabrielse, Rev. Mod. Phys. 58, 233 (1986); L. Lamata, D. Porras, and J. I. Cirac,
Phys. Rev. A 81, 022301 (2010).
[4] P.M.Platzman and M.I. Dykman, Science 284, 1967 (1999).
[5] M. I.Dykman, P.M.Platzman, and P.Seddighrad, Phys. Rev . B67, 155402 (2003).
[6] E.Collin,W.Bailey,P.Fozooni,P.G.Frayne,P.Glasson ,K.Harrabi,M.J.Lea,andG.Papageorgiou,
Phys. Rev. Lett. 89, 245301 (2002).
[7] S.A.Lyon, Phys. Rev.A 74, 052338 (2006).
[8] F. R. Bradbury, M. Takita, T. M. Gurrieri, K. J. Wilkel, K. Eng, M. S. Carroll, and S. A. Lyon, Phys.
Rev. Lett. 107, 266803 (2011); K.Kono, Physics, 4, 110 (2011).
[9] G. Papageorgiou, P. Glasson, K. Harrabi, V. Antonov, E. C ollin, P. Fozooni, P. G. Frayne, M. J. Lea,
and D. G. Rees, Appl. Phys. Lett. 86, 153106 (2005); G. Sabouret, F. R. Bradbury, S. Shankar, J. A .
11Bert, and S.A.Lyon, Appl. Phys. Lett. 92, 082104 (2008).
[10] M. Koch, G.Aub ¨ ock, C.Callegari, and W.E.Ernst, Phys. Rev. Lett. 103, 035302 (2009).
[11] M. Zhang, H.Y. Jiaand L.F.Wei, Phys. Rev.A 80, 055801 (2009).
[12] M. Zhang, H.Y. Jiaand L.F.Wei, Opt. Lett. 351686 (2010).
[13] D. I. Schuster, A. Fragner, M. I. Dykman, S. A. Lyon, and R . J. Schoelkopf, Phys. Rev. Lett. 105,
040503 (2010).
[14] S.Mostame and R.Sch ¨utzhold, Phys. Rev. Lett. 101, 220501 (2008).
[15] C. C. Grimes and T. R. Brown, Phys. Rev. Lett. 32, 280 (1974); D. Konstantinov, M. I. Dykman, M.
J. Lea, Y.Monarkha, and K.Kono, Phys.Rev. Lett. 103, 096801 (2009).
[16] S.S.Sokolov, Phys.Rev. B 51, 2640 (1995).
[17] Alternatively, the single-bit operations could be imp lemented by applying an ac current I(t)(with
frequency ω=νsand phase θon the electrode I2, see, Fig. 2), which generates a magnetic field
Bz≈µ0I(t)d/[2π(d2+h2)]along the z-direction toexcite resonantly the spin of electron e1.
[18] C. Monroe, D. M. Meekhof, B. E. King, W. M. Itano, and D. J. Wineland, Phys. Rev. Lett. 754714
(1995).
[19] K.Mølmer and A.Sørensen, Phys. Rev.Lett. 821835 (1999).
[20] K. R. Brown, C. Ospelkaus, Y. Colombe, A. C. Wilson, D. Le ibfried, and D. J. Wineland, Nature
(London) 471, 196 (2011).
[21] M. Harlander, R.Lechner, M.Brownnutt, R.Blatt, and W. H¨ ansel, Nature (London) 471, 200 (2011).
[22] M. Zhang and L.F.Wei, Phys. Rev.A, 83064301 (2011).
12 |
2012.02810v1.Effects_of_hybridization_and_spin_orbit_coupling_to_induce_odd_frequency_pairing_in_two_band_superconductors.pdf | arXiv:2012.02810v1 [cond-mat.supr-con] 4 Dec 2020✐✐
“paper” — 2020/12/8 — 1:58 — page 1 — #1
✐✐
✐
✐✐
✐Effects of hybridization and spin-orbit
coupling to induce odd frequency pairing in
two-band superconductors
Moloud Tamadonpour and Heshmatollah Yavari
The effects of spin independent hybridization potential and spin-
orbit coupling on two-band superconductor with equal time s -wave
interband pairing order parameter is investigated theoret ically. To
study symmetry classes in two-band superconductors the Gor ’kov
equations are solved analytically. By defining spin singlet and spin
triplet s-wave order parameter due to two-band degree of fre e-
dom the symmetry classes of Cooper pair are studied. For spin
singlet case it is shown that spin independent hybridizatio n gen-
erates Cooper pair belongs to even-frequency spin singlet e ven-
momentum even-band parity (ESEE) symmetry class for both in -
traband and interband pairing correlations. For spin tripl et order
parameter, intraband pairing correlation generates odd-f requency
spin triplet even-momentum even-band parity (OTEE) symme-
try class whereas, interband pairing correlation generate s even-
frequency spin triplet even-momentum odd-band parity (ETE O)
class. For the spin singlet, spin-orbit coupling generates pairing cor-
relation that belongs to odd-frequency spin singlet odd-mo mentum
even-band parity (OSOE) symmetry class and even-frequency spin
singlet even-momentum even-band parity (ESEE) for intraba nd
and interband pairing correlation respectively. In the spi n triplet
case for itraband and interband correlation, spin-orbit co upling
generates even-frequency spin triplet odd-momentum even- band
parity (ETOE) and even-frequency spin triplet even-moment um
odd-band parity (ETEO) respectively.
1. Introduction and summary
Symmetries of order parameter in superconductors affect the ir physical prop-
erties. The total wave function of a pair of fermions, in acco rdance with the
Pauli principle, should be asymmetric under the permutatio n of orbital, spin
and time (or equivalently Matsubara frequency) coordinate s [1]. This leads
1✐✐
“paper” — 2020/12/8 — 1:58 — page 2 — #2
✐✐
✐
✐✐
✐2 Moloud Tamadonpour and Heshmatollah Yavari
to four classes allowed combinations for the symmetries of t he wave function.
This would imply that if the pairing is even in time, spin sing let pairs have
even parity (ESE) and spin triplet pairs have odd parity (ETO ). While if
the pairing is odd in time, spin singlet pairs have odd parity (OSO) and
spin triplet pairs have even parity (OTE). Black-Schaffer and Balatsky [2]
have shown that the multiband superconducting order parame ter has an ex-
tra symmetry classification that originates from the band de gree of freedom,
so called even-band-parity and odd band-parity. As a conseq uence, Cooper
pairs can be classified into eight symmetry classes [3].
Transport properties of multi-band superconductor are qua litatively dif-
ferent from those of the one-band superconductor. For insta nce, two-band
system with the non-magnetic impurity violates Anderson th eorem [4]. As
a result, lots of efforts have been devoted to understanding t he properties
of such systems both theoretically and experimentally. For these materials
band symmetry plays important role. A main hypothesis of the model is the
formation of the Cooper pairs inside one energy band and tran sition of this
pair from one band to another which leads to intra and inter ba nd electronic
interactions. Multi-band model explained lots of strange p hysical properties
of superconductive systems and were consistent with experi mental data. Fa-
mous multiband superconductors are MgB2 [5, 6] and the iron-b ased super-
conductors [7–9]. The nature of their two bands requires tha t the multiband
approach be used to describe their properties. On the contra ry for cuprates
despite their multiband nature a single-band approach is mo re appropriate.
From a general symmetry analysis of even and odd-frequency p airing
states, it was shown that odd-frequency pairing always exis ts in the form of
odd-interband (orbital) pairing if there is any even-frequ ency even-interband
pairing present consistent with the general symmetry requi rements [10]. The
appearance of odd-frequency Cooper pairs in two-band super conductors by
solving the Gor’kov equation was discussed analytically [1 1]. They considered
the equal-time s-wave pair potential and introduced two typ es of hybridiza-
tion potentials between the two conduction bands. One is a sp in-independent
hybridization potential and the other is a spin-dependent h ybridization po-
tential derived from the spin-orbit interaction.
The effect of random nonmagnetic impurities on the supercond ucting
transition temperature in a two-band superconductor, by as suming the equal-
time spin-singlet s-wave pair potential in each conduction band and the hy-
bridization between the two bands as well as the band asymmet ry was stud-
ied theoretically [11, 12]. The effect of single-quasiparti cle hybridization or
scattering in a two-band superconductor by performing pert urbation theory✐✐
“paper” — 2020/12/8 — 1:58 — page 3 — #3
✐✐
✐
✐✐
✐Effects of hybridization and spin-orbit coupling 3
to infinite order in the hybridization term, in a multiband su perconductor
was investigated [13].
The superconducting state of multi-orbital spin-orbit cou pled systems in
the presence of an orbitally driven inversion asymmetry, by assuming that
the interorbital attraction is the dominant pairing channe l, was studied [14].
They have shown that in the absence of the inversion symmetry , supercon-
ducting states that avoid mixing of spin-triplet and spin-s inglet configura-
tions are allowed, and remarkably, spin-triplet states tha t are topologically
nontrivial can be stabilized in a large portion of the phase d iagram. The
impact of strong spin-orbit coupling (SOC) on the propertie s of new class
superconductors has attracted much attentions. It has been the subject of
great theoretical and experimental interest [15, 16]. The f ormation of unex-
pected multi-component superconductors states allows for superconductors
with magnetism and SOC. It was shown that for multi-orbital s ystems such
as the Fe-pnictides SOC coupling, is much smaller than the or bit Hund’s
coupling [17–19], In contrast for multiband systems such as Ir-based oxide
materials it was found that the SOC interaction is comparabl e to the on-site
Coulomb interaction [20]. The combined effect of Hund’s and S OC coupling
on superconductivity in multi-orbital systems was investi gated and it was
shown that Hund’s interaction leads to orbital-singlet spi n-triplet supercon-
ductivity, where the Cooper pair wave function is antisymme tric under the
exchange of two orbitals [21]. Combined effect of the spin-or bit coupling and
scattering on the nonmagnetic disorder on the formation of t he spin reso-
nance peak in iron-based superconductors was also studied [ 22].
In this paper by using Gor’kov equation the effects of spin-or bit coupling
and hybridization on the possibility of odd frequency pairi ng of a two-band
superconductor with an equal time s-wave interband pairing order parameter
are investigated theoretically.
2. Formalism
2.1. Two-band model
The basic physics of multiband superconductors can be obtai ned by intro-
ducing a two-band model. We start with a normal two-band Hami ltonian as✐✐
“paper” — 2020/12/8 — 1:58 — page 4 — #4
✐✐
✐
✐✐
✐4 Moloud Tamadonpour and Heshmatollah Yavari
[12]
(2.1)
ˇHN=/integraldisplay
dr/bracketleftig
ψ†
1,↑(r), ψ†
1,↓(r), ψ†
2,↑(r), ψ†
2,↓(r)/bracketrightig
ˇHN(r)
ψ1,↑(r)
ψ1,↓(r)
ψ2,↑(r)
ψ2,↓(r)
,
where
(2.2) ˇHN=/parenleftbiggξ1kˆσ0/parenleftbig
υeiθ+V/parenrightbig
ˆσ0 /parenleftbig
υe−iθ+V∗/parenrightbig
ˆσ0ξ2kˆσ0/parenrightbigg
.
Hereψα,σ(r)is the annihilation ( ψ†
α,σ(r))creation) operator of an electron
with spin (σ=↑,↓) at theαth conduction band, ξαk=/planckover2pi12k2/2me−µFis the
dispersion energy of band α,meis the mass of an electron, µFis the chemical
potential. The spin independent hybridization potential i s a complex number
characterized by a phase θ.υeiθdenotes the hybridization between the two
bands, which is much smaller than the Fermi energy in the two c onduction
bands. In the absence of spin flip hybridization the spin-orb it coupling poten-
tial isV(k) =ηˆz.(σ×/vectork) =η(kyσx−kxσy), whereηis the parameter that
describes the strength of the Rashba spin-orbit coupling an dˆzis the unit
vector perpendicular to the superconducting surface. This potential is odd-
momentum-parity functions satisfying V(k) =−V(−k). Throughout this pa-
per, Pauli matrices in spin, two-band, particle- hole space s are respectively
denoted by ˆσj,ˆρjandˆτjforj= 1−3. Superconducting order parameter in
bandαis:
(2.3) ˆ∆αα′(k) =/parenleftbigg∆11(k) ∆12(k)
∆21(k) ∆22(k)/parenrightbigg
.
We focus only on interband superconducting order parameter (∆11(k) =
∆22(k) = 0) . The interband s-wave pair potential, is defined by [12]
(2.4) ∆12,σσ′(r) =g/an}bracketle{tψ1,σ(r)ψ2,σ′(r)/an}bracketri}ht.
heregis interband attractive interaction between two electrons . By assuming
the spatially uniform order parameter the Fourier transfor mation of the pair
potential becomes
(2.5) ∆12,↑↓=g
Vvol/summationdisplay
k/an}bracketle{tψ1,↑(k)ψ2,↓(−k)/an}bracketri}ht.✐✐
“paper” — 2020/12/8 — 1:58 — page 5 — #5
✐✐
✐
✐✐
✐Effects of hybridization and spin-orbit coupling 5
In the two-band model, for spin singlet the order parameter i s symmetric
(antisymmetric) under the permutation of band (spin) indic es.
(2.6) ∆12,↑↓= ∆21;↑↓=−∆1,2;↓↑,
But for spin triplet the order parameter is antisymmetric (sy mmetric) under
the permutation of band (spin) indices.
(2.7) ∆12,↑↓=−∆21;↑↓= ∆1,2;↓↑.
For simplicity we omit the indices of ∆αα′. The Hamiltonian describing su-
perconductor in the Nambu space, can be written as [11]
(2.8)⌣HS(T)=1
2/summationdisplay
kψ†
k,σ/parenleftiggˇHN(k)ˇ∆S(T)
ˇ∆†
S(T)−ˇH∗
N(−k)/parenrightigg
ψk,σ,
where the spin-singlet and spin triplet pair potentials ( ˇ∆Sandˇ∆T) are
respectively given by
(2.9) ˇ∆S= ∆ˆρ1iˆσ2,
(2.10) ˇ∆T= ∆iˆρ2ˆσ1.
For a two-band system, the Bogoliubov- de Gennes Hamiltonian can be de-
scribed by 8×8matrix reflecting spin, particle- hole and two band degrees
of freedom. In particle-hole space N1, by considering the spin of electron as
↑and for hole as ↓, while in particle-hole space N2, we consider the spin of
electron as ↓and for hole as ↑, we can describe the Hamiltonian ˇHS(T)by a
4×4matrix [3, 12]
(2.11)
ˇH0=
ξkυeiθ+V(k) 0 ∆
υe−iθ+V∗(k)ξk −sspin∆ 0
0 −sspin∆ −ξk −υe−iθ−V∗(−k)
∆ 0 −υeiθ−V(−k) −ξk
heresspin=−1for spin singlet and sspin= 1for spin triplet. To discuss
the effects of hybridizations and spin-orbit interaction on the properties of
superconductors, we calculate the Green’s functions by sol ving the Gor’kov✐✐
“paper” — 2020/12/8 — 1:58 — page 6 — #6
✐✐
✐
✐✐
✐6 Moloud Tamadonpour and Heshmatollah Yavari
equation [23]
(2.12)/parenleftbig
iωn−ˇH0/parenrightbigˇG0(k,iωn) =ˇ1,
(2.13) ˇG0(k,iωn) =/parenleftigg
ˆG0(k,iωn) ˆF0(k,iωn)
−sspinˆF†
0(−k,iωn)−ˆG∗
0(−k,iωn)/parenrightigg
.
whereωn= (2n+1)πkBTis the Matsubara frequency ( kBis the Boltz-
mann constant), and ˇ1is the identity matrix in spin×band×particle−
holespace.ˇG0is a4×4matrix where the diagonal components are nor-
mal Green’s function and non-diagonal components are anoma lous Green’s
function.
2.2. Spin Singlet Pairing Order
According to Equation (2.11), the Hamiltonian of a two-band superconductor
with spin singlet configuration in the presence of spin-orbi t coupling is
(2.14)
ˇH0=
ξk υeiθ+η(ky+ikx) 0 ∆
υe−iθ+η(ky−ikx) ξk ∆ 0
0 ∆ −ξk −υe−iθ+η(ky−ikx)
∆ 0 −υeiθ+η(ky+ikx) −ξk
.
By using Equation (2.12) and (2.13) , for spin singlet the solu tion of the
normal Green’s function within the first order of ∆is calculated as
(2.15)
ˆG0(k,iωn) =∆
Z0{[(ξ−iωn)(ν2+η2k2−2νη(kxsinθ+kycosθ)+(ξ+iωn)
×/parenleftbig
−ξ2−ω2
n/parenrightbig
]ˆρ0+[(−νcosθ−ηky)(−(ξ+iωn)2+ν2+η2k2
−2νη(kxsinθ+kycosθ))]ˆρ1+[(νsinθ+ηkx)(−(ξ+iωn)2
+ν2+η2k2−2νη(kxsinθ+kycosθ))]ˆρ2}
here
(2.16)
Z0=ξ4+2ξ2/parenleftbig
ω2
n−ν2/parenrightbig
+/parenleftbig
ω2
n+ν2/parenrightbig2−8iηνξωn(kxsinθ+kycosθ)
+2cos2θη2ν2(k2
x−k2
y)−4sin2θη2ν2kxky+2η2k2/parenleftbig
ω2
n−ξ2/parenrightbig
+η4k4.
thatkx=kcosφandky=ksinφ, whereφis the angle between momentum
and thexaxis. The matrix form of the normal Green’s function ( Eq. (2. 15))✐✐
“paper” — 2020/12/8 — 1:58 — page 7 — #7
✐✐
✐
✐✐
✐Effects of hybridization and spin-orbit coupling 7
can be written as
(2.17) ˆG0(k,iωn) =/parenleftbiggG11(k,iωn)G12(k,iωn)
G21(k,iωn)G22(k,iωn)/parenrightbigg
.
where
(2.18)
G11(k,iωn) =∆
Z0[(ξ−iωn)(ν2+η2(k2
x+k2
y)−2νη(kxsinθ+kycosθ)
−(ξ+iωn)(ξ2+ω2
n)]
(2.19)
G12(k,iωn) =∆
Z0[{−νeiθ−η(ikx+ky)}{−(ξ+iωn)2+ν2+η2k2
−2νη(kxsinθ+kycosθ)}]
(2.20)
G21(k,iωn) =∆
Z0[{−νe−iθ−η(−ikx+ky)}{−(ξ+iωn)2+ν2+η2k2
−2νη(kxsinθ+kycosθ)}]
(2.21)
G22(k,iωn) =∆
Z0[(ξ−iωn)(νe−iθ−η(−ikx+ky))(νeiθ−η(ikx+ky)
−(ξ+iωn)/parenleftbig
ξ2+ω2
n/parenrightbig
]
By using Equation (2.12) and (2.13), the anomalous Green’s fu nction can be
obtained as
(2.22)
ˆF0(k,iωn) =∆
Z0[(2νξcosθ+2ηωniky) ˆρ0+/parenleftbig
−(ν2+ξ2+ω2
n)+η2k2/parenrightbig
ˆρ1
+(2νη(kxcosθ−kysinθ)) ˆρ2+(2νξisinθ−2ηωnkx) ˆρ3].
In particle- hole space N1, the matrix form of the anomalous Green’s func-
tion (Eq. (2.22)) is
(2.23) ˆFN1
0(k,iωn) =/parenleftbiggF11,↑↓(k,iωn)F12,↑↓(k,iωn)
F21,↑↓(k,iωn)F22,↑↓(k,iωn)/parenrightbigg
,
where
(2.24) F11,↑↓(k,iωn) =∆
Z0[2νξeiθ−2ηωnkx+2ηωniky],
(2.25)
F12,↑↓(k,iωn) =∆
Z0[−(ν2+ξ2+ω2
n)+η2k2−2iνη(kxcosθ−kysinθ)],✐✐
“paper” — 2020/12/8 — 1:58 — page 8 — #8
✐✐
✐
✐✐
✐8 Moloud Tamadonpour and Heshmatollah Yavari
(2.26)
F21,↑↓(k,iωn) =∆
Z0[−(ν2+ξ2+ω2
n)+η2k2+2iνη(kxcosθ−kysinθ)],
(2.27) F22,↑↓(k,iωn) =∆
Z0[2νξeiθ−2ηωnkx+2ηωniky].
In particle- hole space N2, the matrix form of the anomalous Green’s function
is
(2.28)ˆFN2
0(k,iωn) =/parenleftbigg
F11,↓↑(k,iωn)F12,↓↑(k,iωn)
F21,↓↑(k,iωn)F22,↓↑(k,iωn)/parenrightbigg
=−ˆFN1
0(k,iωn).
In the absence of spin-orbit coupling ( η) the anomalous Green’s function
(Eq.(2.22)) becomes
(2.29) ˆF0(k,iωn) =∆
Z0[2νξcosθˆρ0−(ν2+ξ2+ω2
n)ˆρ1+2νξisinθˆρ3],
here
(2.30) Z0=ξ4+2ξ2/parenleftbig
ω2
n−ν2/parenrightbig
+/parenleftbig
ω2
n+ν2/parenrightbig2.
The matrix form of the anomalous Green’s function in Equatio n (2.29) in
particle-hole spaces N1andN2, are
(2.31)
ˆFN1
0(k,iωn) =/parenleftbigg
F11,↑↓(k,iωn)F12,↑↓(k,iωn)
F21,↑↓(k,iωn)F22,↑↓(k,iωn)/parenrightbigg
=∆
Z0/parenleftbigg
2νξcosθ+2iνξsinθ−(ν2+ξ2+ω2
n)
−(ν2+ξ2+ω2
n) 2νξcosθ−2iνξsinθ/parenrightbigg
=∆
Z0/parenleftbigg
2ξνeiθ−(ν2+ξ2+ω2
n)
−(ν2+ξ2+ω2
n) 2ξνe−iθ/parenrightbigg
,
and
(2.32)
ˆFN2
0(k,iωn) =/parenleftbiggF11,↓↑(k,iωn)F12,↓↑(k,iωn)
F21,↓↑(k,iωn)F22,↓↑(k,iωn)/parenrightbigg
=−ˆFN1
0(k,iωn)
=∆
Z0/parenleftbigg
−2νξcosθ−2iνξsinθ(ν2+ξ2+ω2
n)
(ν2+ξ2+ω2
n)−2νξcosθ+2iνξsinθ/parenrightbigg
=∆
Z0/parenleftbigg−2ξνeiθ(ν2+ξ2+ω2
n)
(ν2+ξ2+ω2
n)−2ξνe−iθ/parenrightbigg
.✐✐
“paper” — 2020/12/8 — 1:58 — page 9 — #9
✐✐
✐
✐✐
✐Effects of hybridization and spin-orbit coupling 9
The intraband pairing correlations become
(2.33) F11,↑↓(k,iωn)−F11,↓↑(k,iωn) =4∆
Z0ξνeiθ,
(2.34) F22,↑↓(k,iωn)−F22,↓↑(k,iωn) =4∆
Z0ξνe−iθ.
Hybridization generates ρ0andρ3components which belongs to even fre-
quency symmetry class. It means that in the presence of inter band cou-
pling, hybridization generates even frequency intra- subl attice pairing in
the system. These components belong to even-frequency spin -singlet even-
momentum even-band parity (ESEE) symmetry class. This resu lt is in agree-
ment with the equation (20) presented in Ref [12]. Equation ( 2.33) and (2.34)
are in agreement with the Equation (62) and (63) reported in R ef [3] in the
first order of ∆(|∆|2= 0) and equal energy bands ( ξ−= 0) and both belong
to the (ESEE) symmetry class. The band symmetry generates in terband
pairing correlation:
(2.35)[F12,↑↓(k,iωn)−F12,↓↑(k,iωn)]+[F21,↑↓(k,iωn)−F21,↓↑(k,iωn)]
=−4∆
Z0(ν2+ξ2+ω2
n).
which belongs to (ESEE). This result is in agreement with the Equation (65)
presented in Ref [3].
(2.36)[F12,↑↓(k,iωn)−F12,↓↑(k,iωn)]−[F21,↑↓(k,iωn)−F21,↓↑(k,iωn)]
=2∆
Z3iωnξ−.
which belongs to the symmetry (OSEO) class. We considered a t wo-band
superconductor with an equal dispersion energy in each band (ξ+=ξ−). In
this case the interband pairing correlation due to band asym metry is
(2.37)
[F12,↑↓(k,iωn)−F12,↓↑(k,iωn)]−[F21,↑↓(k,iωn)−F21,↓↑(k,iωn)] = 0.
In the absence of hybridization within the second order of th e spin-orbit
coupling constant ( η), we obtain
(2.38)
ˆG0(k,iωn) =∆
Z0[(η2k2−(ξ+iωn)2)(ξ−iωn) ˆρ0+η(ky+ikx)(ξ+iωn)2ˆρ1],
where
(2.39) Z0= (ξ2+ω2
n)2+2η2k2/parenleftbig
ω2
n−ξ2/parenrightbig
.✐✐
“paper” — 2020/12/8 — 1:58 — page 10 — #10
✐✐
✐
✐✐
✐10 Moloud Tamadonpour and Heshmatollah Yavari
Equation (2.22) can be rewritten as
(2.40) ˆF0(k,iω) =∆
Z0[2iηωnkyˆρ0−2ηωnkxˆρ3+/parenleftbig
−ξ2−ω2
n+η2k2/parenrightbig
ˆρ1].
In particle- hole space N1andN2,the matrix form of the anomalous Green’s
function (Eq. (2.40)) is
(2.41)ˆFN1
0(k,iωn) =/parenleftbigg
F11,↑↓(k,iωn)F12,↑↓(k,iωn)
F21,↑↓(k,iωn)F22,↑↓(k,iωn)/parenrightbigg
=∆
Z0/parenleftbigg2iηωnky−2ηωnkx−ξ2−ω2
n+η2k2
−ξ2−ω2
n+η2k22iηωnky+2ηωnkx/parenrightbigg
,
(2.42)
ˆFN2
0(k,iωn) =/parenleftbiggF11,↓↑(k,iωn)F12,↓↑(k,iωn)
F21,↓↑(k,iωn)F22,↓↑(k,iωn)/parenrightbigg
=−ˆFN1
0(k,iωn)
=∆
Z0/parenleftbigg−2iηωnky+2ηωnkxξ2+ω2
n−η2k2
ξ2+ω2
n−η2k2−2iηωnky−2ηωnkx/parenrightbigg
.
The intraband pairing correlations are
(2.43) F11,↑↓(k,iωn)−F11,↓↑(k,iωn) =4∆
Z0iωnη(ky+ikx),
(2.44) F22,↑↓(k,iωn)−F22,↓↑(k,iωn) =4∆
Z0iωnη(ky−ikx).
Spin-orbit coupling generates ρ0andρ3components which belong to odd
frequency symmetry class. It means that in the presence of in ter-band cou-
pling, spin-orbit coupling generates odd-frequency intra sublattice pairing in
the system. These components belong to odd-frequency spin- singlet odd-
momentum even-band parity (OSOE) symmetry class. In Ref [3] the intra-
band pairing correlation is written as
(2.45) F11,↑↓(k,iωn)+F11,↓↑(k,iωn) =∆
Z3(ξ+−ξ−)V3,
(2.46) F22,↑↓(k,iωn)+F22,↓↑(k,iωn) =∆
Z3(ξ++ξ−)V3.✐✐
“paper” — 2020/12/8 — 1:58 — page 11 — #11
✐✐
✐
✐✐
✐Effects of hybridization and spin-orbit coupling 11
The hybridization generates pairing correlations that bel ong to the (ETOE)
class. The band asymmetry generates interband pairing corr elation as
(2.47)
[F12,↑↓(k,iωn)−F12,↓↑(k,iωn)]−[F21,↑↓(k,iωn)−F21,↓↑(k,iωn)] = 0.
In Ref [3] for spin orbit hybridization the band asymmetry ge nerates the
interband pair correlation as
(2.48)
[F12,↑↓(k,iω)−F12,↓↑(k,iω)]−[F21,↑↓(k,iω)−F21,↓↑(k,iω)] =2∆
Z3iωnξ−.
which belongs to the odd-frequency spin-singlet even-mome ntum odd-band
parity symmetry (OSEO). The interband pairing correlation due to band
symmetry is
(2.49)[F12,↑↓(k,iωn)−F12,↓↑(k,iωn)]+[F21,↑↓(k,iωn)−F21,↓↑(k,iωn)]
=−4∆
Z0(ξ2+ω2
n−η2k2).
This component belongs to even-frequency spin-singlet eve n-momentum even-
band parity (ESEE) symmetry class. For spin singlet, hybrid ization potential
generates ESEE symmetry class due to both intra and interban d correlation,
whereas the spin dependent hybridization potential genera tes this class only
for interband pairing correlation due to band symmetry. In t his case the odd
frequency pairing arises only due to intraband pairing corr elations for spin
dependent hybridization potential.
2.3. Spin Triplet Pairing Order
By considering Equation (2.11), the Hamiltonian of a two-ban d supercon-
ductor with spin triplet configuration in the presence of spi n-orbit coupling
is
(2.50)
ˇH0=
ξk υeiθ+η(ky+ikx) 0 ∆
υe−iθ+η(ky−ikx) ξk −∆ 0
0 −∆ −ξk −υe−iθ+η(ky−ikx)
∆ 0 −υeiθ+η(ky+ikx) −ξk
.
The solution of the anomalous Green’s function within the fir st order of ∆
is calculated as
(2.51)
ˆF0(k,iωn) =∆
Z0[(−2iηξkx+2νωnsinθ)ˆρ0+(2iνη(kxcosθ−kysinθ)) ˆρ1
+(i(ν2−ξ2−ω2
n)−iη2(k2
x+k2
y))ˆρ2+(−2ηξky−2iνωncosθ)ˆρ3].✐✐
“paper” — 2020/12/8 — 1:58 — page 12 — #12
✐✐
✐
✐✐
✐12 Moloud Tamadonpour and Heshmatollah Yavari
The matrix form of the anomalous Green’s function (Eq. (2.51 ) ) can be
written as
(2.52) ˆF11,↑↓(k,iωn) =∆
Z0/parenleftig
−2iνωneiθ−2iηξkx−2ηξky/parenrightig
,
(2.53)
ˆF12,↑↓(k,iωn) =∆
Z0[(ν2−ξ2−ω2
n)−η2k2+2iνη(kxcosθ−kysinθ)],
(2.54)
ˆF21,↑↓(k,iωn) =∆
Z0[(−ν2+ξ2+ω2
n)+η2k2+2iνη(kxcosθ−kysinθ)],
(2.55) ˆF22,↑↓(k,iωn) =∆
Z0/parenleftig
2iνωne−iθ−2iηξkx+2ηξky/parenrightig
.
In the absence of spin-orbit coupling ( η= 0) the anomalous Green’s
function Equation (2.51) becomes
(2.56)ˆF0(k,iωn) =∆
Z0[2νωnsinθˆρ0+i(ν2−ξ2−ω2
n)ˆρ2−2iνωncosθˆρ3].
here
(2.57) Z0=ξ4+2ξ2/parenleftbig
ω2
n−ν2/parenrightbig
+/parenleftbig
ω2
n+ν2/parenrightbig2.
In particle-hole space N1andN2, the matrix form of the anomalous Green’s
function (Eq. (2.56)) is
(2.58)
ˆFN1
0(k,iωn) =/parenleftbiggF11,↑↓(k,iωn)F12,↑↓(k,iωn)
F21,↑↓(k,iωn)F22,↑↓(k,iωn)/parenrightbigg
=∆
Z0/parenleftbigg
2νωn(sinθ−icosθ) (ν2−ξ2−ω2
n)
−(ν2−ξ2−ω2
n) 2νωn(sinθ+icosθ)/parenrightbigg
,
(2.59)ˆFN2
0(k,iωn) =/parenleftbiggF11,↓↑(k,iωn)F12,↓↑(k,iωn)
F21,↓↑(k,iωn)F22,↓↑(k,iωn)/parenrightbigg
=ˆFN1
0(k,iωn)
=∆
Z0/parenleftbigg
2νωn(sinθ−icosθ) (ν2−ξ2−ω2
n)
−(ν2−ξ2−ω2
n) 2νωn(sinθ+icosθ)/parenrightbigg
.
The intraband pairing correlations becomes
(2.60) F11,↑↓(k,iωn)+F11,↓↑(k,iωn) =−4∆
Z0iωnνe−iθ,✐✐
“paper” — 2020/12/8 — 1:58 — page 13 — #13
✐✐
✐
✐✐
✐Effects of hybridization and spin-orbit coupling 13
(2.61) F22,↑↓(k,iωn)+F22,↓↑(k,iωn) =4∆
Z0iωnνeiθ.
Hybridization generates ρ0andρ3which belongs to odd frequency sym-
metry class. These components belong to odd-frequency spin -triplet even-
momentum even-band parity (OTEE) symmetry class. This resu lt is in agree-
ment with the Equation (24) presents in Ref [12] In the first or der of∆
(|∆|2= 0) and equal energy bands ( ξ−= 0) Equation (2.60) and (2.61)
are coincide with the Equation (83) and (84) presented in Ref [3] and both
belong to the (OTEE) symmetry class. The band symmetry gener ates inter-
band pairing correlation as
(2.62)[F12,↑↓(k,iωn)+F12,↓↑(k,iωn)]−[F21,↑↓(k,iωn)+F21,↓↑(k,iωn)]
=4∆
Z0(ν2−ξ2−ω2
n).
which belongs to even-frequency spin triplet even-momentu m odd-band par-
ity (ETEO) symmetry class. In Ref [3] the interband pairing c orrelation due
to band asymmetry is
(2.63)[F12,↑↓(k,iωn)+F12,↓↑(k,iωn)]+[F21,↑↓(k,iωn)+F21,↓↑(k,iωn)]
=2∆
Z5iωnξ−.
Thus the band hybridization generates pairing correlation s that belong to
the odd-frequency spin triplet even-momentum even-band pa rity (OTEE)
class. Since we considered a two-band superconductor with a n equal energy
bands, the interband pairing correlation due to band asymme try is
(2.64)
[F12,↑↓(k,iωn)+F12,↓↑(k,iωn)]+[F21,↑↓(k,iωn)+F21,↓↑(k,iωn)] = 0.
In the absence of hybridization, we obtain
(2.65) ˆF0(k,iωn) =∆
Z0[−2iηξkxˆρ0−2ηξkyˆρ3−i/parenleftbig
ξ2+ω2
n+η2k2/parenrightbig
ˆρ2].
The matrix form of the anomalous Green’s function in Equatio n (2.65) in
particle- hole spaces N1andN2, are
(2.66)ˆFN1
0(k,iωn) =/parenleftbiggF11,↑↓(k,iωn)F12,↑↓(k,iωn)
F21,↑↓(k,iωn)F22,↑↓(k,iωn)/parenrightbigg
=∆
Z0/parenleftbigg−2iηξkx−2ηξky−(ξ2+ω2
n+η2k2)
(ξ2+ω2
n+η2k2)−2iηξkx+2ηξky/parenrightbigg
,✐✐
“paper” — 2020/12/8 — 1:58 — page 14 — #14
✐✐
✐
✐✐
✐14 Moloud Tamadonpour and Heshmatollah Yavari
(2.67)ˆFN2
0(k,iωn) =/parenleftbigg
F11,↓↑(k,iωn)F12,↓↑(k,iωn)
F21,↓↑(k,iωn)F22,↓↑(k,iωn)/parenrightbigg
=ˆFN1
0(k,iωn)
=∆
Z0/parenleftbigg
−2iηξkx−2ηξky−(ξ2+ω2
n+η2k2)
(ξ2+ω2
n+η2k2)−2iηξkx+2ηξky/parenrightbigg
.
The intraband pairing correlations are
(2.68) F11,↑↓(k,iωn)+F11,↓↑(k,iωn) =−4∆
Z0ξη(ky+ikx),
(2.69) F22,↑↓(k,iωn)+F22,↓↑(k,iωn) =4∆
Z0ξ(ky−ikx).
Spin-orbit coupling generates ˆρ0andˆρ3which belong to even frequency sym-
metry class. These components belong to even-frequency spi n-triplet odd-
momentum even-band parity (ETOE) symmetry class. In Ref [3] the intra-
band pairing correlation is calculated as
(2.70) F11,↑↓(k,iωn)−F11,↓↑(k,iωn) =−∆
Z5iωnV3,
(2.71) F22,↑↓(k,iωn)−F22,↓↑(k,iωn) =∆
Z0iωnV3.
The hybridization generates pairing correlations that bel ong to the odd-
frequency spin singlet odd-momentum even-band parity (OSO E) class. As
mentioned in Ref [3] the interband pair correlation can be wr itten as
(2.72)[F12,↑↓(k,iωn)+F12,↓↑(k,iωn)]+[F21,↑↓(k,iωn)+F21,↓↑(k,iωn)]
=2∆
Z5iωnξ−.
Thus the spin-orbit coupling generates pairing correlatio ns that belong to
the odd-frequency spin triplet even-momentum even-band pa rity (OTEE)
class. In contrast in our formalism the band asymmetry gener ates interband
pairing correlation as
(2.73)
[F12,↑↓(k,iωn)+F12,↓↑(k,iωn)]+[F21,↑↓(k,iωn)+F21,↓↑(k,iωn)] = 0.
The interband pairing correlation due to band symmetry is
(2.74)[F12,↑↓(k,iωn)+F12,↓↑(k,iωn)]−[F21,↑↓(k,iωn)+F21,↓↑(k,iωn)]
=−4∆
Z0(ξ2+ω2
n+η2k2).✐✐
“paper” — 2020/12/8 — 1:58 — page 15 — #15
✐✐
✐
✐✐
✐Effects of hybridization and spin-orbit coupling 15
These components belong to even-frequency spin-triplet ev en-momentum
odd-band parity (ETEO) symmetry class. Thus, for spin tripl et, the spin
dependent and spin independent hybridization both generat e the same sym-
metry class ETEO due to interband pairing correlation. The o dd frequency
pairing arises in the presence of spin independent hybridiz ation due to in-
traband pairing correlations.
3. Conclusion
Within the theoretical model the existence of odd frequency pairs in two
band superconductors by incorporating both spin independe nt hybridization
and spin dependent spin-orbit interaction is investigated . This model also
includes both the one-particle hybridization term and all p ossible intraband
and interband superconducting pairing interaction terms i n a two-band sys-
tem.
The normal and anomalous thermal Green’s functions have bee n calcu-
lated in the Nambu formalism as elements of the Fourier trans formed4×4
matrix Green’s function by taking into account of all possib le intraband
and interband superconducting interaction terms coupling both bands in
the mean field approximation. By assuming that the attractive interaction
acts on two electrons with different spins in different conduc tion bands dif-
ferent symmetry classes were demonstrated in the presence o f hybridization
and spin-orbit coupling.
The role of intraband and interband pairing correlations to emerge the
odd frequency in a two-band superconductor was examined. Fo r spin singlet,
the odd-frequency is generated by spin dependent hybridiza tion potential
owing to intraband pairing correlations in agreement with t he odd frequency
generated by the interband pair correlation due to band asym metry in Ref
[3]. On the other hand, for spin triplet the spin independent hybridization
potential generates the odd-frequency pairing due to intra band correlations
in agreement with the result of Ref [12].
References
[1] M. Sigrist and K. Ueda, Phenomenological theory of unconventional su-
perconductivity . Rev. Mod. Phys., 63, 239, (1991).
[2] A. M. Black-Schaffer and A. V. Balasky, Proximity-induced unconven-
tional superconductivity in topological insulators . Phys. Rev. B, 87,
220506(R), (2013).✐✐
“paper” — 2020/12/8 — 1:58 — page 16 — #16
✐✐
✐
✐✐
✐16 Moloud Tamadonpour and Heshmatollah Yavari
[3] Y. Asano and A. Sasaki, Odd-frequency Cooper pairs in two-band su-
perconductors and their magnetic response . Phys. Rev. B, 92, 224508,
(2015).
[4] V. A. Moskalenko and M. E. Palistrant, Two-band model determination
of the critical temperature of a superconductor with an impu rity. Sov.
Phys. JETP, 22, 526, (1966).
[5] X. X. Xi, Two-band superconductor magnesium diboride . Rep. Prog.
Phys., 71, 116501, (2008).
[6] J. Nagamatsu, N. Nakagawa, T. Muranaka, Y. Zenitani, J. A kimutsu,
Superconductivity at 39 K in magnesium diboride . Nature, 410, 63,
(2001).
[7] Y. Shun-Li and L. Jian-Xin, Spin fluctuations and unconventional su-
perconducting pairing in iron-based superconductors . Chin. Phys. B, 22,
087411, (2013).
[8] Y. Tanaka, P. M. Shirage and A. Iyo, Disappearance of Meissner Effect
and Specific Heat Jump in a Multiband Superconductor . J Supercond
Nov Magn, 23, 253–256, (2010).
[9] P. O. Sprau et al, Discovery of orbital-selective Cooper pairing in FeSe .
science, 357, 75-80, (2013).
[10] A. M. Black-Schaffer and A. V. Balasky, Odd-frequency superconducting
pairing in multiband superconductors . Phys. Rev. B, 88, 104514, (2013).
[11] Y. Asano and A. A. Golubov, Green’s-function theory of dirty two-band
superconductivity . Phys. Rev. B, 97, 214508, (2018).
[12] Y. Asano, A Sasaki and A. A. Golubov, Dirty two-band superconductivity
with interband pairing order . New J. Phys., 20, 043020, (2018).
[13] L. Komendov’a, A. V. Balatsky, and A. M. Black-Schaffer, Band hy-
bridization induced odd-frequency pairing in multiband su perconductors .
Phys. Rev. B, 92, 094517, (2015).
[14] Y. Fukaya, S. Tamura, K. Yada K, Y. Tanaka, P. Gentile, an d M. Cuoco,
Interorbital topological superconductivity in spin-orbi t coupled supercon-
ductors with inversion symmetry breaking . Phys. Rev. B , 97, 174522,
(2018).
[15] V. P. Mineev and M. Sigrist, Basic theory of superconductivity in metals
without inversion center . Lect. Notes Phys., 847, 129–154, (2012).✐✐
“paper” — 2020/12/8 — 1:58 — page 17 — #17
✐✐
✐
✐✐
✐Effects of hybridization and spin-orbit coupling 17
[16] M. Smidman, M. B Salamon, H. Q. Yuan and D. F. Agterberg, Spin-
triplet p-wave pairing in a three-orbital model for iron pni ctide super-
conductors . Rep. Prog. Phys, 80, 036501, (2017).
[17] P. Lee and X. G. Wen, S, Superconductivity and spin-orbit coupling in
non-centrosymmetric materials: A review . EPhys. Rev. B, 78, 2, (2008).
[18] W. L. Yang et al, Evidence for weak electronic correlations in iron pnic-
tidesPhys. Rev. B, 80, 014508, (2009).
[19] P. Fazekas, Lecture Notes on Electron Correlation and Magnetism .
World Scientific, Singapore, (1999).
[20] H. Kuriyama et al, Epitaxially stabilized iridium espinel oxide without
cations in the tetrahedral site . Appl. Phys. Lett., 96, 27010, (2010).
[21] M. P. Christoph and H. Y. Kee, Identifying spin-triplet pairing in spin-
orbit coupled multi-band superconductors . EPL, 98, 2, (2011).
[22] M. M. Korshunov and Y. N. Togushova, Spin-orbit coupling and im-
purity scattering on the spin resonance peak in three orbita l model for
Fe-based superconductors . Journal of Siberian Federal University Math-
ematics and Physics, 11, 998, (2018).
[23] L. P. Gor’kov, Theory of superconducting alloys in a strong magnetic
field near the critical tepmerature . JETP, 10, 998, (1960).
Department of Physics, University of Isfahan,
Isfahan 81746, Iran
E-mail address :h.yavary@sci.ui.ac.ir✐✐
“paper” — 2020/12/8 — 1:58 — page 18 — #18
✐✐
✐
✐✐
✐ |
1205.2162v3.3D_quaternionic_condensations__Hopf_invariants__and_skyrmion_lattices_with_synthetic_spin_orbit_coupling.pdf | arXiv:1205.2162v3 [cond-mat.quant-gas] 10 Feb 20163D quaternionic condensations, Hopf invariants, and skyrm ion lattices with synthetic
spin-orbit coupling
Yi Li,1,2Xiangfa Zhou,3and Congjun Wu1
1Department of Physics, University of California, San Diego , La Jolla, California 92093, USA
2Princeton Center for Theoretical Science, Princeton Unive rsity, Princeton, NJ 08544
3Key Laboratory of Quantum Information, University of Scien ce and Technology of China, CAS, Hefei, Anhui 230026, China
We study the topological configurations of the two-componen t condensates of bosons with the
3D/vector σ·/vector pWeyl-type spin-orbit coupling subject to a harmonic trappi ng potential. The topology
of the condensate wavefunctions manifests in the quaternio nic representation. In comparison to
theU(1) complex phase, the quaternionic phase manifold is S3and the spin orientations form
theS2Bloch sphere through the 1st Hopf mapping. The spatial distr ibutions of the quaternionic
phases exhibit the 3D skyrmion configurations, and the spin d istributions possess non-trivial Hopf
invariants. Spin textures evolve from the concentric distr ibutions at the weak spin-orbit coupling
regime to the rotation symmetry breaking patterns at the int ermediate spin-orbit coupling regime.
In the strong spin-orbit coupling regime, the single-parti cle spectra exhibit the Landau-level type
quantization. In this regime, the three-dimensional skyrm ion lattice structures are formed when
interactions are below the energy scale of Landau level mixi ngs. Sufficiently strong interactions
can change condensates into spin-polarized plane-wave sta tes, or, superpositions of two plane-waves
exhibiting helical spin spirals.
PACS numbers: 03.75.Mn, 03.75.Lm, 03.75.Nt, 67.85.Fg
I. INTRODUCTION
Quantum mechanical wavefunctions generally speak-
ing are complex-valued. However, for the single com-
ponent boson systems, their ground state many-body
wavefunctions are highly constrained, which are usu-
ally positive-definite [1], as a consequence of the Perron-
Frobeniustheoreminthemathematicalcontextofmatrix
analysis [2]. This is a generalization of the “no-node”
theorem of the single-particle quantum mechanics, for
example, both the ground state wavefunctions of har-
monic oscillators and hydrogen atoms are nodeless. Al-
though the positive-definiteness does not apply to the
many-body fermion wavefunctions because Fermi statis-
tics necessarilyleads to nodal structures, it remains valid
for many-body boson systems. It applies under the fol-
lowing conditions: the Laplacian type kinetic energy, the
arbitrary single-particle potential, and the coordinate-
dependent interactions. The positive-definiteness of the
ground state wavefunctions implies that time-reversal
(TR) symmetry cannot be spontaneously broken in con-
ventional Bose-Einstein condensates (BEC), such as the
superfluid4He and most ground state BECs of ultra-cold
alkali bosons [3].
It would be interesting to seek unconventional BECs
beyond the constraint of positive-definite condensate
wavefunctions [4]. The spin-orbit coupled boson systems
areanideal platformto studythis classofexoticstatesof
bosons, which can spontaneously breaking the TR sym-
metry. In addition to a simple Laplacian, the kinetic
energy contains the spin-orbit coupling term linearly de-
pendent on momentum. If the bare interaction is spin-
independent, the condensate wavefunctions are heavily
degenerate. An “order-from-disorder” calculation based
on the zero-point energy of the Bogoliubov spectra wasperform to select the condensate configuration[4]. Inside
the harmonic trap, it is predicted that the condensates
spontaneously develop the half-quantum vortex coexist-
ing with 2D skyrmion-type spin textures [5]. Experi-
mentally, spin-orbit coupled bosons have been realized
in exciton systems in semi-conducting quantum wells.
Spin texture configurations similar to those predicted in
Ref. [5] have been observed [6]. On the other hand,
the progress of synthetic artificial gauge fields in ultra-
cold atomic gases greatly stimulates the investigation of
the above exotic states of bosons [7, 8]. Extensive stud-
ies have been performed for bosons with the 2D Rashba
spin-orbit coupling, which exhibit various spin structures
arising from the competitions among the spin-orbit cou-
pling, interaction,andtheconfiningtrapenergy[5,9–16].
Most studies so far have been on the two-dimensional
spin-orbit coupled bosons. It would be interesting to fur-
ther consider the unconventional condensates of bosons
with the three-dimensional Weyl-type spin-orbit cou-
pling, whose experimental realization has been proposed
by the authors through atom-light interactions in a com-
bined tripod and tetrapod level system [20] and also by
Anderson et al.[21]. As will be shown below, the qua-
terinon representationprovides a natural and most beau-
tiful description of the topological condensation configu-
rations. Quaternions are an extension of complex num-
bers as the first discovered non-commutative division al-
gebra, which has provided a new formulation of quantum
mechanics [17–19]. Similarly to complex numbers whose
phasesspanaunitcircle S1, thequaternionicphasesspan
a three dimensional unit sphere S3. The spin distribu-
tions associated with quaternionic wavefunctions are ob-
tained through the 1st Hopf map S3→S2as will be
explained below. It would be interesting to search for
BECswith non-trivialtopologicaldefects associatedwith2
the quaternionic phase structure. It will be a new class
of unconventional BECs beyond the “no-node” theorem
breaking TR symmetry spontaneously.
In this article, we consider the unconventional conden-
satewavefunctionswiththe3DWeyl-typespin-orbitcou-
pling/vector σ·/vector p. The condensation wavefunctions exhibit topo-
logically non-trivial configurations as 3D skyrmions, and
spin density distributions are also non-trivial with non-
zero Hopf invariants. These topological configurations
can be best represented as defects of quaternion phase
distributions. Spatial distributions of the quaternionic
phase textures and spin textures are concentric at weak
spin-orbit couplings. As increasing spin-orbit coupling,
these textures evolve to lattice structures which are the
3D quaternionic analogy of the 2D Abrikosov lattice of
the usual complex condensate.
The rest part of this article is organized as follows.
In Sect. II, we define the model Hamiltonian. In Sect.
III, the condensate wavefunctions in the weak spin-orbit
regimearestudied. Topologicalanalysesonthe skyrmion
configurations and Hopf invariants are performed by us-
ing the quaternion representation. In Sect. IV, the
skyrmion lattice configuration of the spin textures is
studied in the intermediate and strong spin-orbit cou-
pling regimes. In Sect. V, superpositions of plane-wave
condensate configurations are studied. Conclusions are
made in Sect. VI.
II. THE MODEL HAMILTONIAN
We consider a two-component boson system with the
3D spin-orbit coupling of the /vector σ·/vector p-type confined in a har-
monic trap. The free part of the Hamiltonian is defined
as
H0=/integraldisplay
d3/vector r ψ†
γ(/vector r)/braceleftBig
−/planckover2pi12/vector∇2
2m+i/planckover2pi1λ/vector σγδ·(/vector∇)
+1
2mω2/vector r2/bracerightBig
ψδ(/vector r), (1)
whereγandδequal↑and↓referring to two internal
states of bosons; /vector σare Pauli matrices; mis the bo-
son mass;λis the spin-orbit coupling strength with the
unit of velocity; ωis the trap frequency. At the single-
particle level, Eq. (1) satisfies the Kramer-type time-
reversal symmetry of T= (−iσ2)Cwith the property of
T2=−1. However, parity is broken by spin-orbit cou-
pling. In the absence ofthe trap, good quantum numbers
for the single-particle states are the eigenvalues ±1 of he-
licity/vector σ·/vector p/|p|, wherepis the momentum. This results in
two branches of dispersions
ǫ±(/vectork) =/planckover2pi12
2m(k∓kso)2, (2)
where/planckover2pi1kso=mλ. The lowest single-particle energy
states lie in the sphere with the radius ksodenoted as the
spin-orbit sphere. It corresponds to a spin-orbit lengthscalelso= 1/ksoin real space. The harmonic trap has a
natural length scale lT=/radicalBig
/planckover2pi1
mω, and thus the dimension-
less parameter α=lTksodescribes the relative spin-orbit
coupling strength.
As for the interaction Hamiltonian, we use the contact
s-wave scattering interaction defined as
Hint=gγδ
2/integraldisplay
d3/vector r ψ†
γ(/vector r)ψ†
δ(/vector r)ψδ(/vector r)ψγ(/vector r).(3)
Two different interaction parameters are allowed, in-
cluding the intra and inter-component ones defined as
g↑↑=g↓↓=g, andg↑↓=cg, wherecis a constant.
In the previous study of the 2D Rashba spin-orbit cou-
pling with harmonic potentials [5, 15], the single-particle
eigenstates are intuitively expressed in the momentum
representation: the low energy state lies around a ring in
momentum space, and the harmonic potential becomes
the planar rotor operator on this ring subject to a π-
flux, which quantizes the angular momentum jzto half
integers. Similar picture also applies in 3D [5, 22]. The
low energy states are around the spin-orbit sphere. In
the projected low energy Hilbert space, the eigenvectors
read
ψ+(/vectork) = (cosθk
2,sinθk
2eiφk)T. (4)
The harmonic potential is again a rotor Hamiltonian on
the spin-orbit sphere subject to the Berry gauge connec-
tion as
Vtp=1
2m(i∇k−/vectorAk)2(5)
with the moment of inertial I=Mkk2
soandMk=
/planckover2pi12/(mω2)./vectorAk=i/angb∇acketleftψ+(/vectork)|∇k|ψ+(/vectork)/angb∇acket∇ightis the vector po-
tential of a U(1) magnetic monopole, which quantizes
the angular momentum jto half-integers. While the ra-
dial energy is still quantized in terms of /planckover2pi1ω, the angular
energydispersion with respect to jis stronglysuppressed
at large values of αas
Enr,j,jz≈/parenleftBig
nr+j(j+1)
2α2/parenrightBig
/planckover2pi1ω+const,(6)
wherenris the radial quantum number. As further
shown in Ref. [20], in the case α≫1, all the states
with the same nrbut different jandjzare nearly degen-
erate, thus can be viewed as one 3D Landau level with
spherical symmetry but the broken parity. If filled with
fermions, the system belongs to the Z2-class of 3D strong
topological insulators.
Now we load the system with bosons. The interaction
energy scale is defined as Eint=gN0/l3
T, whereN0is
the total particle number in the condensate. The corre-
sponding dimensionless parameter is β=Eint//planckover2pi1ω. At
the Hartree-Fock level, the Gross-Pitaevskii energy func-
tional is defined in terms of the condensate wavefunction3
Ψ = (Ψ ↑,Ψ↓)Tas
E=/integraldisplay
d3/vector r(Ψ†
↑,Ψ†
↓)/braceleftBig
−/planckover2pi12∇2
2m−iλ/planckover2pi1/vector∇·/vector σ+1
2mω2r2
+g/parenleftbigg
n↑+cn↓0
0cn↑+n↓/parenrightbigg/bracerightBig/parenleftbigg
Ψ↑
Ψ↓/parenrightbigg
, (7)
wheren↑,↓(/vector r) =N0|Ψ↑,↓(/vector r)|2are the particle densities of
two components, respectively, and Ψ( /vector r) is normalized as/integraltext
d3/vector rΨ†(/vector r)Ψ(/vector r) = 1. The condensate wavefunction Ψ( /vector r)
is solved numerically by using the standard method of
imaginary time evolution. The dimensionless form of the
Gross-Pitaevskii equation is
E′=/integraldisplay
d3/vector r′(˜Ψ†
↑,˜Ψ†
↓)/braceleftBig
−/vector∇′2
2−iα/vector∇′·/vector σ+r′2
2
+β/parenleftbigg
˜n↑+c˜n↓0
0c˜n↑+ ˜n↓/parenrightbigg/bracerightBig/parenleftbigg˜Ψ↑
˜Ψ↓/parenrightbigg
,(8)
whereE′=E/(/planckover2pi1ω),/vector∇′=lT/vector∇;/vector r′=/vector r/lT;˜Ψ↑and˜Ψ↓
arethe renormalizedcondensate wavefunctionssatisfying/integraltext
d3r′|˜Ψ↑|2+|˜Ψ↓|2= 1; ˜n↑=|˜Ψ↑|2and ˜n↓=|˜Ψ↓|2.
III. THE WEAK SPIN-ORBIT COUPLING
REGIME
In this section, we consider the condensate configura-
tion in the limit of weak spin-orbit coupling, say, α∼1.
In this regime, the single-particle spectra still resemble
those of the harmonictrap. We study the case that inter-
actions arenot strongenough to mix stateswith different
angular momenta.
A. The spin-orbit coupled condensate
In this regime, the condensate wavefunction Ψ remains
the same symmetry structure as the single-particle wave-
function over a wide range of interaction parameter β,
i.e., Ψ remains the eigenstates of j=1
2as confirmed
numerically below. Ψ can be represented as
Ψj=jz=1
2(r,ˆΩ) =f(r)Y+
j,jz(ˆΩ)+ig(r)Y−
j,jz(ˆΩ),(9)
wheref(r) andg(r) are real radial functions. Y±
j,jz(ˆΩ)
are the spin-orbit coupled spherical harmonic functions
with even and odd parities, respectively. For example,
for the case of j=jz=1
2, they are
Y+
1
2,1
2(r,ˆΩ) =/parenleftbigg
1
0/parenrightbigg
, Y−
1
2,1
2(r,ˆΩ) =/parenleftbiggcosθ
sinθeiφ/parenrightbigg
,(10)
whose orbital partial-wavecomponents are sandp-wave,
respectively. The TR partner of Eq. (9) is ψjz=−1
2=
ˆTψj=jz=1
2=iσ2ψ∗
j=jz=1
2. The two terms in Eq. (9)
are of opposite parity eigenvalues, mixed by the paritybreaking spin-orbit coupling /vector σ·/vector p. The coefficient iof the
Y−
jjzterm is because the matrix element /angb∇acketleftY+
jjz|/vector σ·/vector p|Y−
jjz/angb∇acket∇ight
is purely imaginary.
For the non-interacting case, the radial wavefunctions
uptoaGaussianfactorcanbeapproximatedbyspherical
Bessel functions as
f(r)≈j0(ksor)e−r2/2l2
T, g(r)≈j1(ksor)e−r2/2l2
T,(11)
which correspond to the sandp-partial waves, respec-
tively. Both of them oscillate along the radial direction
and the pitch values are around kso. Atr= 0,f(r)
reaches the maximum and g(r) is 0. As rincreases,
roughly speaking, the zero points of f(r) corresponds to
the extrema of g(r) and vise versa. Repulsive interac-
tions expand the spatial distributions of f(r) andg(r),
but the above picture still holds qualitatively. In other
words, there is aπ
2-phase shift between the oscillations
off(r) andg(r).
B. The quaternion representation
Can we have unconventional BECs with non-trivial
quaternionic condensate wavefunctions? Actually, the
topological structure of condensate wavefunction Eq. (9)
manifests clearly in the quaternion representation as
shown below.
We define the following mapping from the complex
two-component vector Ψ = (Ψ ↑,Ψ↓)Tto a quaternion
variable through
ξ=ξ0+ξ1i+ξ2j+ξ3k, (12)
where
ξ0= ReΨ ↑,ξ1= ImΨ ↓,ξ2=−ReΨ↓,ξ3= ImΨ ↑.(13)
i,j,kare the imaginary units satisfying i2=j2=k2=
−1, and the anti-commutation relation ij=−ji=k.
The TR transformation on ξis just−jξ.
Eq. (9) can be expressed in the quaternionic exponen-
tial form as
ξj=jz=1
2(r,ˆΩ) =|ξ(r)|e/vector ω(ˆΩ)γ(r)=|ξ|(cosγ+/vector ωsinγ),(14)
where
|ξ(r)|= [f2(r)+g2(r)]1
2,
/vector ω(ˆΩ) = sinθcosφ i+sinθsinφ j+cosθ k,
cosγ(r) =f(r)/|ξ(r)|,sinγ(r) =g(r)/|ξ(r)|.(15)
ω(ˆΩ) is the imaginary unit along the direction of ˆΩ sat-
isfying/vector ω2(ˆΩ) =−1. According to the oscillating proper-
ties off(r) andg(r),γ(r) spirals as rincreases. At the
n-th zero point of g(r) denotedrn,γ(rn) =nπwhere
n≥0 and we define r0= 0, while at the n-th zero point
off(r) denotedr′
n,γ(r′
n) = (n−1
2)πwheren≥1.
In 3D, the condensate wavefunctions can be topolog-
ically non-trivial because the homotopy group of the4
quaternionic phase is π3(S3) =Z[23, 24]. The corre-
sponding winding number, i.e. the Pontryagin index, of
the mapping S3→S3is the 3D skyrmion number. The
spatial distribution of the quaternionic phase e/vector ω(ˆΩ)γ(r)
defined in Eq. 14, which lies on S3, exhibits a topo-
logically nontrivial mapping from R3toS3, i.e., a 3D
multiple skyrmion configuration. This type of topolog-
ical defects are non-singular which is different from the
usual vortex in single component BEC. In realistic trap-
ping systems, the coordinate space is the open R3. At
large distance r≫lT,|ξ(r)|decays exponentially, where
the quaternionic phase and the mapping are not well-defined. Nevertheless, in each concentric spherical shell
withrn<r<r n+1,γ(r) winds from nπto (n+1)π, and
ω(ˆΩ) covers all the directions, thus this shell contributes
1 to the winding number of e/vector ω(ˆΩ)γ(r)onS3. If the system
size is truncated at the order of lT, the skyrmion number
can be approximated at the order of lTkso=α.
There exists an interesting difference from the previ-
ously studied 2D case: Although the spin density dis-
tribution exhibit the 2D skyrmion configuration due to
π2(S2) [5, 15, 16], the 2D condensation wavefunctions
have no well-defined topology due to π2(S3) = 0.
FIG. 1: The distribution of /vectorS(/vector r) in a) the xz-plane and in the horizontal planes with b)z= 0 andc)z/lT=1
2.
The unit length is set as lT= 1 in all the figures in this article. The color scale shows the magnitude of out-plane
component Syina) andSzinb) andc). The parameter values are α= 1.5,c= 1, andβ= 30, and the length unit
in these and all the figures below is lT.
C. The Hopf mapping and Hopf invariant
Exotic spin textures in spinor condensates have been
extensively investigated [25–27]. In our case, the 3D spin
density distributions /vectorS(/vector r) exhibit a novel configuration
with non-trivial Hopf invariants due to the non-trivial
homotopy group π3(S2) =Z[23, 24]./vectorS(/vector r) can be ob-
tained from ξ(r) through the 1st Hopf map defined as
/vectorS(/vector r) =1
2ψ†
γ/vector σγβψδ, or, in the quaternionicrepresentation,
1
2¯ξkξ=Sxi+Syj+Szk, (16)
where¯ξ=ξ0−ξ1i−ξ2j−ξ3kisthequaternionicconjugate
ofξ. The Hopf invariant of the 1st Hopf map is just 1
[24]. The real space concentric spherical shell rn< r <
rn+1maps to the quaternionic phase S3, and the latter
furthermapstothe S2Blochspherethroughthe1stHopf
map. The winding number of the first map is 1, and the
Hopf invariant of the second map is also 1, thus the Hopf
invariantoftheshell rn<r<r n+1toS2is1. Rigorously
speaking, the magnitude of /vectorS(/vector r) decays exponentially at
r≫lT, and thus the total Hopf invariant is not well-
defined in the open R3space. Again, if we truncate thesystem size at lT, the Hopf invariant is approximately at
the order of α.
FIG. 2: The Hopf fibration of the spin texture configura-
tion in Fig. 1. Every circle represents a spin orientation,
and every two circles are linked with the linking number
1.
Next we present numeric results for the spin textures
associated with the condensation wavefunction Eq. 9 as5
plotted in Fig. 1. Explicitly, /vectorS(/vector r) is expressed as
/bracketleftbigg
Sx(/vector r)
Sy(/vector r)/bracketrightbigg
=g(r)sinθ/bracketleftbigg
cosφ−sinφ
sinφcosφ/bracketrightbigg/bracketleftbigg
g(r)cosθ
f(r)/bracketrightbigg
,
Sz(/vector r) =f2(r)+g2(r)cos2θ, (17)
In thexz-plane, the in-plane components SxandSzform
a vortex in the half plane of x >0 andSyis prominent
in the core. The contribution at large distance is ne-
glected, where /vectorS(/vector r) decays exponentially. Due to the
axial symmetry of /vectorS(/vector r) in Eq. 17, the 3D distribution is
just a rotation of that in Fig. 1 a) around the z-axis. In
thexy-plane, spin distribution exhibits a 2D skyrmion
pattern, whose in-plane components are along the tan-
gential direction. As the horizontal cross-section shifted
along thez-axis,/vectorS(/vector r) remains 2D skyrmion-like, but its
in-plane components are twisted around the z-axis. The
spin configuration at z=−z0can be obtained by a com-bined operation of TR and rotation around the y-axis
180◦, thus its in-plane components are twisted in an op-
posite way compared to those at z=z0. Combining
the configurations on the vertical and horizontal cross
sections, we complete the 3D distribution of /vectorS(/vector r) with
non-zero Hopf invariant.
The non-trivial structure of the Hopf invariant of the
above spin configuration can be revealed by plotting its
Hopf fibration in terms of the linked non-crossing circles
in real space, as shown in Fig. 2. For all the points on
each circle, their normalized spin polarizations /angb∇acketleft/vector σ/angb∇acket∇ight/|/angb∇acketleft/vector σ/angb∇acket∇ight|
are the same, corresponding to a single point on the S2
sphere. Inaddition, everytwocirclesarelinkedwith each
other with the linking number 1, which is the standard
Hopf bundle structure describing a many-to-one map
fromS3toS2. Ultracoldbosonswithsyntheticspin-orbit
coupling providea novel platform to study such beautiful
mathematical ideas in realistic physics systems.
FIG. 3: The distribution of /vectorS(/vector r) in horizontal cross-sections with a) z/lT=−0.5, b)z/lT= 0, c)z/lT= 0.5,
respectively. The color scale shows the value of Sz, and parameter values are α= 4,β= 2, andc= 1.
IV. THE INTERMEDIATE AND STRONG
SPIN-ORBIT COUPLING REGIME
A. The intermediate spin-orbit coupling strength
Next we consider the case of the intermediate spin-
orbit coupling strength, i.e., 1 < α <10, at which the
single-particle spectra evolve from the case of the har-
monic potential to Landau level-like as shown in Eq. 6.
Interactions are sufficiently strong to mix a few lowest
energy states with different angular momenta j. As a re-
sult, rotationalsymmetryisbrokenandcomplexpatterns
appear.
In this case, the topology of condensate wavefunctions
isstill 3Dskyrmion-likemapping from R3toS3, andspin
textures with the non-trivial Hopf invariant are obtained
through the 1st Hopf map. Compared to the weak spin-
orbit coupling case, the quaternionic phase skyrmions
and spin textures are no longer concentric, but split to a
multi-centered pattern. The numeric results of /vectorS(/vector r) areplotted in Fig. 3 for different horizontal cross-sections.
In thexy-plane,/vectorSexhibits the 2D skyrmion pattern as
shown in Fig. 3 ( b): The in-plane components form two
vortices and one anti-vortex, while Sz’s inside the vortex
and anti-vortex cores are opposite in direction, thus they
contribute to the skyrmion number with the same sign.
The spin configuration at z=z0>0 is shown in Fig. 3
(a), which is twisted around the z-axis clock-wise. After
performing the combined TR and rotation around the y-
axis 180◦, we arrive at the configuration at z=−z0in
Fig. 3(b).
B. The strong spin-orbit coupling regime
We next consider the case of strong spin-orbit cou-
pling, i.e., α≫1. The single-particle spectra already
exhibit the Landau-level type quantization in this regime
as shown in Eq. 6. The single-particle eigenstates with
nr= 0 are nearly degenerate i.e., they form the low-6
est Landau level states. We assume that the interaction
strength is enough to mix states inside the lowest Lan-
dau level but is still relatively weak not to induce inter-
Landau level mixing.
In this regime, the length scale of each skyrmion is
shortened as enlarging the spin-orbit coupling strength.
As we can imagine, more and more skyrmions appear
and will form a 3D lattice structure, which is the SU(2)
generalization of the 2D Abrikosov lattice of the usual
U(1) superfluid. We have numerically solved the Gross-
PitaevskiiequationEq. 7andfound thelattice structure:
Each lattice site is a single skyrmion of the condensate
wavefunction ξ(/vector r), whose spin configuration exhibits the
texture configuration approximately with a unit Hopf in-
variant. The numeric results for the spin texture config-uration are depicted in Fig. 4 a) andb) for two different
horizontal cross sections parallel to the xy-plane. In each
cross section, spin textures form a square lattice, and the
lattice constant dis estimated approximately the spin-
orbit length scale as
d≃2πlso= 2πlT/α. (18)
For two horizontal cross sections with a distance of
∆z≃d/2, their square lattice configurations are dis-
placedalongthediagonaldirection: Thesitesatonelayer
sit abovethe plaquette centersofthe adjacentlayer. As a
result, the overall three-dimensional configuration of the
topological defects is a body-centered cubic ( bcc) lattice,
and its size is finite confined by the trap.
FIG. 4: The distribution of /vectorS(/vector r) in horizontal cross-sections with (a) z/lT= 0, (b)z/lT= 0.2, respectively. The color
scale shows the value of Szand parameter values are α= 22,β= 1, andc= 1. The overall lattice exhibits the bcc
structure.
V. THE EFFECT OF STRONG INTERACTIONS
In this section, we present the condensate configura-
tions in the case that both spin-orbit coupling and inter-
actions are strong, such that different Landau levels are
mixed by interactions.
In this case, the effect of the harmonic trapping poten-
tial becomes weak compared with interaction energies,
thus we can approximate the condensate wavefunctions
as superpositions of plane-wave states. The plane-wave
components are located on the spin-orbit sphere and the
condensate wavefunctions are no longer topological. At
c= 1, the interaction is spin-independent, and bosons
select a superposition of a pair of states ±/vectorkon the spin-
orbit sphere, say, ±ksoˆz. The condensate wavefunction
is written as
ψ(/vector r) =/radicalbigg
Na
N0eiksoz| ↑/angb∇acket∇ight+/radicalbigg
Nb
N0e−iksoz| ↓/angb∇acket∇ight,(19)withNa+Nb=N0. The density of Eq. 19 in real
space is uniform to minimize the interaction energy at
the Hartree-Fock level. However, all the different parti-
tions ofNa,byield the same Hartree-Fock energy. The
quantum zero point energy from the Bogoliubov modes
removes this accidental degeneracy through the “order-
from-disorder” mechanism, which selects the equal par-
titionNa=Nb. The calculation is in parallel to that
of the 2D Rashba case performed in Ref. [5], thus will
not be presented here. In this case, the condensate is a
spin helix propagates along z-axis and spin spirals in the
xy-plane.
Atc/negationslash= 1, the spin-dependent part of the interaction
can be written as
Hsp=1−c
2g/integraldisplay
d3r(ψ†
↑ψ↑−ψ†
↓ψ↓)2.(20)
Atc >1, the interaction energy at Hartree-Fock level
is minimized for the condensate wavefunction of a plane
wave state eiksoz| ↑/angb∇acket∇ight, or, its TR partner.7
Forc<1,/angb∇acketleftHsp/angb∇acket∇ightis minimized if /angb∇acketleftSz/angb∇acket∇ight= 0 in space. At
the Hartree-Fock level, the condensate can either be a
plane-wave state with momentum lying in the equator of
the spin-orbit sphere and spin polarizing in the xy-plane,
or, the spin spiral state described by Eq. 19 with Na=
Nb. An“order-from-disorder”analysisontheBogoliubov
zero-point energies indicates that the spin spiral stateis selected. We also present the numerical results for
Eq. (4) in the main text with a harmonic trap in Fig.
5 for the case of c <1. The condensate momenta of
two spin components have opposite signs, thus the trap
inhomogeneity already prefers the spin spiral state Eq.
19 at the Hartree-Fock level.
FIG. 5: The density profile (a) for ↑-component, and that for ↓-component is the same. Phase profiles for (b) ↑and
(c)↓- components, respectively. Parameter values are a= 10,β= 50, andc= 0.5.
VI. CONCLUSION
In summary, we have investigated the two-component
unconventional BECs driven by the 3D spin-orbit
coupling. In the quaternionic representation, the
quaternionic phase distributions exhibit non-trivial 3D
skyrmionconfigurationsfrom R3toS3. Thespinorienta-
tion distributions exhibit texture configurations charac-
terizedby non-zeroHopfinvariantsfrom R3toS2. These
two topological structures are connected through the 1st
Hopf map from S3toS2. At large spin-orbit coupling
strength, the crystalline order of spin textures, or, wave-
function skyrmions, are formed, which can be viewed as
a generalization of the Abrikosov lattice in 3D.
Note added.— Near the completion of this manuscript,
we became aware of a related work by Kawakami et al.[28], in which the condensate wavefunction in the weak
spin-orbit coupling case was studied.
Acknowledgments.— Y.L. thanks the Princeton Cen-
ter for Theoretical Science at Princeton University for
support. X. F. Z. acknowledges the support of NFRP
(2011CB921204, 2011CBA00200), the Strategic Priority
Research Program of the Chinese Academy of Sciences
(Grant No. XDB01030000),NSFC (11004186,11474266),
and the Major Research plan of the National Natural
Science Foundation of China (91536219). C. W. is sup-
ported by the NSF DMR-1410375 and AFOSR FA9550-
14-1-0168. C. W. acknowledges the support from the
Presidents Research Catalyst Awards of University of
California, and National Natural Science Foundation of
China (11328403).
[1] R. P. Feynman, Statistical Mechanics, A Set of Lectures
(Berlin: Addison-Wesley, 1972).
[2] R. B. Bapat and T. Raghavan, Non-Negative Matri-
ces and Applications (Cambridge University Press, Cam-
bridge, United Kingdom, 1997).
[3] A. J. Leggett, Rev. Mod. Phys. 73, 307 (2001).
[4] C. Wu, Mod. Phys. Lett. 23, 1 (2009).
[5] C. Wu , I. Mondragon-Shem, arXiv:0809.3532; C. Wu ,
I. Mondragon-Shem, and X. F. Zhou, Chin. Phys. Lett.,
28, 097102 (2011).
[6] A.A. High et al., Nature 483, 584 (2012). A.A. High et
al., arXiv:1103.0321.
[7] Y.-J. Lin et al., Nature 462, 628 (2009).[8] Y.-J. Lin et al., Nature 471, 83 (2011).
[9] T. Stanescu et al., Phys. Rev. A 78, 023616 (2008).
[10] T.-L. Ho et al., Phys. Rev. Lett. 107, 150403 (2011).
[11] C. Wang et al., Phys. Rev. Lett. 105, 160403 (2010).
[12] S.-K. Yip, Phys. Rev. A 83, 043616 (2011).
[13] Y. Zhang et al., Phys. Rev. Lett. 108, 035302 (2012).
[14] X.-F. Zhou et al., Phys. Rev. A 84, 063624 (2011).
[15] H. Hu et al., Phys. Rev. Lett. 108, 010402(2012).
[16] S. Sinha, R. Nath, and L. Santos, arXiv:1109.2045.
[17] A. V. Balatsky, cond-mat/9205006.
[18] S. L. Adler, Quaternionic Quantum Mechanics and
Quantum Fields (OxfordUniversityPress, Oxford, 1995).
[19] D.Finkelstein et al., J.Math.Phys.(N.Y.) 3, 207(1962).8
[20] Y. Li et al., Phys. Rev. B 85, 125122 (2012).
[21] B. M. Anderson et al., arXiv:1112.6022.
[22] S. K. Ghosh et al., Phys. Rev. A, 84, 053629 (2011).
[23] F. Wilczek and A. Zee, Phys. Rev. Lett. 51, 2250 (1983).
[24] M. Nakahara, Geometry, topology, and physics , (Taylor
& Francis, 2003)
[25] F. Zhou, Int. J. Mod. Phys. B 17, 2643-2698 (2003); E.Demler and F. Zhou, Phys. Rev. Lett. 88, 163001 (2002).
G. W.SemenoffandF.Zhou, Phys.Rev.Lett. 98, 100401
(2007).
[26] J. Zhang and T. L. Ho, arXiv:0908.1593.
[27] D. M. Stamper-Kurn, and M. Ueda, arXiv:1205.1888.
[28] T. Kawakami et al., arXiv:1204.3177. |
1004.3066v1.Spin_Orbit_Coupling_and_Spin_Waves_in_Ultrathin_Ferromagnets__The_Spin_Wave_Rashba_Effect.pdf | arXiv:1004.3066v1 [cond-mat.mes-hall] 18 Apr 2010Spin Orbit Coupling and Spin Waves in Ultrathin Ferromagnet s:
The Spin Wave Rashba Effect
A. T. Costa,1R. B. Muniz,1S. Lounis,2A. B. Klautau,3and D. L. Mills2
1Instituto de F´ ısica, Universidade Federal Fluminense, 24 210-340 Niter´ oi, RJ, Brasil.
2Department of Physics and Astronomy,
University of California Irvine, California, 92697, U. S. A .
3Departamento de Fisica, Universidade Federal do Par´ a, Bel ´ em, PA, Brazil.
Abstract
We present theoretical studies of the influence of spin orbit coupling on the spin wave excitations
of the Fe monolayer and bilayer on the W(110) surface. The Dzy aloshinskii-Moriya interaction is
active in such films, by virtue of the absence of reflection sym metry in the plane of the film. When
the magnetization is in plane, this leads to a linear term in t he spin wave dispersion relation for
propagation across the magnetization. The dispersion rela tion thus assumes a form similar to
that of an energy band of an electron trapped on a semiconduct or surfaces with Rashba coupling
active. We also show SPEELS response functions that illustr ate the role of spin orbit coupling in
such measurements. In addition to the modifications of the di spersion relations for spin waves, the
presence of spin orbit coupling in the W substrate leads to a s ubstantial increase in the linewidth
of the spin wave modes. The formalism we have developed appli es to a wide range of systems,
and the particular system explored in the numerical calcula tions provides us with an illustration
of phenomena which will be present in other ultrathin ferrom agnet/substrate combinations.
1I. INTRODUCTION
Thestudyofspindynamicsinultrathinferromagnetsisoffundamen talinterest, sincenew
physics arises in these materials that has no counterpart in bulk mag netism. Examples are
provided by relaxation mechanisms evident in ferromagnetic resona nce and Brillouin light
scattering studies,1–3and also for the large wave vectors probed by spin polarized electro n
loss spectroscopy (SPEELS).4Of course, by now the remarkable impact of ultrathin film
structures on magnetic data storage is very well known, and othe r applications that exploit
spin dynamics in such materials are envisioned. Thus these issues are important from a
practical point of view as well as from that of fundamental physics .
Theoretical studies of the nature of spin waves in ultrathin films ads orbed on metal sub-
strates have been carried out for some years now, along with comp arison with descriptions
provided with the Heisenberg model.5In this paper, we extend the earlier theoretical treat-
ments to include the influence of spin orbit coupling on the spin wave sp ectrum of ultrathin
films. This extension is motivated by a most interesting discussion of t he ground state of the
Mn monolayer on the W(110) surface. A nonrelativistic theoretical study of this system pre-
dicted that the ground state would be antiferromagnetic in charac ter.7This prediction was
confirmed by spin polarized scanning tunneling microscope studies of the system.8However,
recent experimental STM data with a more sensitive instrument sho wed a more complex
ground state, wherein the ground state is in fact a spin density wav e.9One can construct
the new state by beginning with the antiferromagnet, and then sup erimposing on this a
long wavelength modulation on the direction of the moments on the lat tice. The authors
of ref. 9 argued that the lack of reflection symmetry of the syste m in the plane of the film
activates theDzyaloshinskii Moriya (DM) interaction, andthe news tate hasitsorigininthis
interaction. They also presented relativistic and ab initio calculations that gave an excellent
account of the new data. The reflection symmetry is broken simply b y the presence of the
substrate upon which the film is grown. This argument to us is most int riguing, since one
can then conclude that the DM interaction must be active in any ultra thin ferromagnet;
the substrate is surely always present. The DM interaction has its o rigin in the spin orbit
interaction, which of course is generally very weak in magnets that in corporate the 3d tran-
sition elements as the moment bearing entities. However, in the case of the Mn monolayer
on W(110) hybridization between the Mn 3d and the W 5d orbitals activ ates the very large
2W spin orbit coupling, with the consequence that the strength of th e DM interaction can
be substantial, as illustrated by the calculations presented in ref. 9 . One may expect to see
substantial impact of the DMinteraction in other ultrathin magnets grown on 5d substrates,
and possibly 4d substrates as well.
We have here another example of new physics present in ultrathin ma gnets that is not
encountered inthebulk formofthematerial fromwhich theultrath instructure isfabricated.
The purpose of this paper is to present our theoretical studies of spin orbit effects on spin
waves and also on the dynamic susceptibility of a much studied ultrath in film/substrate
combination, theFemonolayer andbilayer onW(110). Wefindstriking effects. Forinstance,
when the magnetization is in plane, as we shall see the DM interaction in troduces a term
linear in wave vector in the dispersion relation of spin waves. Thus the uniform spin wave
mode at zero wave vector acquires a finite group velocity. We find th is to be in the range
of 2×105cm/sec for the Fe monolayer on W(110). Furthermore, left/right asymmetries
appear in the SPEELS response functions. Thus, we shall see that spin orbit coupling has
clear effects on the spin excitations of transition metal ultrathin fe rromagnets grown on 5d
substrates.
We comment briefly on the philosophy of the approach used here, an d in various earlier
publications.5Numerous authors proceed as follows. One may generate a descrip tion of the
magnetic ground state of the adsorbed films by means of an electro nic structure calculation
based on density functional theory. It is then possible to calculate , within the framework of
anadiabaticapproximation, effectiveHeisenbergexchangeintegra lsJijbetweenthemagnetic
moments in unit cell i and unit cell j. These may be entered into a Heise nberg Hamilto-
nian, and then spin wave dispersion relations may be calculated throu gh use of spin wave
theory. It has been known for decades10that in the itinerant 3d magnets, effective exchange
interactions calculated in such a manner have very long range in real space. Thus, one must
include a very large number of distant neighbors in order to obtain co nverged results. This
is very demanding to do with high accuracy for the very numerous dis tant neighbors, since
the exchange interactions become very small as one moves out into distant neighbor shells.
At a more fundamental level, as noted briefly above, discussions in e arlier publications
show that in systems such as we study here, the adiabatic approxim ation breaks down badly,
with qualitative consequences.5First, spin wave modes of finite wave vector have very short
lifetimes, by virtue of decay into the continuum of particle hole pairs ( Stoner excitations)
3even at the absolute zero of temperature5,11whereas in Heisenberg model descriptions their
lifetime is infinite. In multi layer films, the earlier calculations show that as a consequence of
the short lifetime, the spectrum of spin fluctuations at large wave v ectors contains a single
broad feature which disperses with wave vector in a manner similar to that of a spin wave;
this is consistent with SPEELS data on an eight layer film of Co on Cu(10 0).4This picture
stands in contrast to that offered by the Heisenberg model, in which a film of N layers has
N spin wave modes for each wave vector, and each mode has infinite lif etime.
The method developed earlier, and extended here to incorporate s pin orbit coupling,
takes due account of the breakdown of the adiabatic approximatio n and also circumvents
the need to calculate effective exchange interactions in real space out to distant neighbor
shells. Weworkdirectlyinwave vectorspacethroughstudyofthew ave vectorandfrequency
dependent susceptibility discussed below, denotedas χ+,−(/vectorQ/bardbl,Ω;l⊥,l′
⊥). Theimaginarypart
of this object, evaluated for l⊥=l′
⊥and considered as a function of frequency Ω for fixed
wave vector /vectorQ/bardblprovides us with the frequency spectrum of spin fluctuations on lay erl⊥for
the wave vector chosen. Spin waves appear as peaks in this functio n, very much as they do
in SPEELS data, and in a manner very similar to that used by experimen talist we extract
a dispersion relation for spin waves by following the wave vector depe ndence of the peak
frequency. We never need to resort to a real space summation pr ocedure over large number
of neighbors, coupled by very tiny exchange couplings. The spin wav e exchange stiffness
can be extracted either by fitting the small wave vector limit of the d ispersion relation so
determined, or alternatively by utilizing an expression derived earlier5which once againdoes
not require a summation in real space.
Wecommentonanotherfeatureofthepresentstudy. Inearlierc alculations,5,11,14asinthe
present paper, an empirical tight binding description forms the bas is for our description of
the electronic structure. Within this approach, referred to as a m ulti band Hubbard model,
we can generate the wave vector and frequency dependent susc eptibility for large systems.
In the earlier papers, effective tight binding parameters were extr acted from bulk electronic
structure calculations. The present studies are based on tight bin ding parameters obtained
directly from a RS-LMTO-ASA calculation for the Fe/W(110) system . We also obtain tight
binding parameters by fitting KKR based electronic structure calcu lations for the ultrathin
film/substrate combinations of interest. We find that spin waves in t he Fe/W(110) system
are quite sensitive to the empirical tight binding parameters which ar e employed, though as
4we shall see the various descriptions provide very similar pictures of the one electron local
density of states.
We note that Udvardi and Szunyogh12have also discussed the influence of spin orbit
coupling on the dispersion relation of spin waves in the Fe monolayer on W(110) within
the framework of the adiabatic approach discussed above, where exchange interactions and
other magnetic parameters are calculated in real space. We shall d iscuss a comparison with
our results and theirs below. There are differences. Most particula rly, we note that in
Fig. 3, the authors of ref. 12 provide two dispersion curves for pr opagation perpendicular
to the magnetization, whereas in a film such as this with one spin per un it cell there can
be only one magnon branch. Additionally and very recently, Bergman n and coworkers45
investigatedwithinanadiabaticapproachfinitetemperatureeffect sonthemagnonspectrum
of Fe/W(110).
Insection II, we comment onourmeans of introducing spin orbit cou pling into the theory.
The results of our calculations are summarized in Section III and con cluding remarks are
found in section IV.
II. CALCULATION OF THE DYNAMIC SUSCEPTIBILITY IN THE PRESENCE
OF SPIN ORBIT COUPLING
Theformalism forincluding spin orbitcoupling effects in our description of spin dynamics
is quite involved, so in this section we confine our attention to an outlin e of the key steps,
and an exposition of the overall structure of the theory. Our sta rting point is the multi band
Hubbard model of the system that was employed in our earlier study of spin dynamics in
ultrathin ferromagnets. The starting Hamiltonian is written as5
H=/summationdisplay
ij/summationdisplay
µνσTµν
ijc†
iµσcjνσ+1
2/summationdisplay
µνµ′ν′/summationdisplay
iσσ′Ui;µν,µ′ν′c†
iµσc†
iνσ′ciν′σ′ciµ′σ (1)
whereiandjaresite indices, σ,σ′refer to spin, and µ,νtothe tight binding orbitals, nine in
numberforeachsite, whichareincludedinourtreatment. TheCoulo mbinteractionsoperate
only within the 3d orbitals on a given lattice site. The film, within which fer romagnetism
is driven by the Coulomb interactions, sits on a semi-infinite substrat e within which the
Coulomb interaction is ignored.
5In our empirical tight binding picture, the spin orbit interaction adds a term we write as
HSO=/summationdisplay
i/summationdisplay
µνλi
2/bracketleftBig
Lz
µν(c†
iµ↑ciν↑−c†
iµciν↓)+L+
µνc†
iµ↓ciν↑+L−
µνc†
iµ↑ciν↓/bracketrightBig
(2)
where/vectorLis the angular momentum operator, λiis the local spin-orbit coupling constant,
L±=Lx±iLyandLα
µν=/an}bracketle{tµ|Lα|ν/an}bracketri}ht. We assume that the spin orbit interaction, present both
within the ferromagnetic film and the substrate, operates only with in the 3d atomic orbitals.
A convenient tabulation of matrix elements of the orbital angular mo mentum operators is
found in ref. 13.
Information on the spin waves follows from the study of the spectr al density of the trans-
verse dynamic susceptibility χ+,−(/vectorQ/bardbl,Ω;l⊥,l′
⊥) as discussed above. From the text around
Eq. (1) of ref. 14, we see that this function describes the amplitud e of the transverse spin
motion (the expectation value of the spin operator S+in the layer labeled l⊥) to a fictitious
transverse magnetic field of frequency Ω and wave vector /vectorQ/bardblparallel to the film surface that
is applied to layer l′
⊥of the sample. The spectral density, given by Im {χ+,−(/vectorQ/bardbl,Ω;l⊥,l′
⊥)},
when multiplied by the Bose Einstein function n(Ω) = [exp( βΩ)−1]−1is also the amplitude
of thermal spin fluctuations of wave vector /vectorQ/bardbland frequency Ω in layer l⊥. We obtain in-
formation regarding the character (frequency, linewidth, and am plitude in layer l⊥) of spin
waves from the study of this function, as discussed earlier.5
Our previous analyses are based on the study of the dynamic susce ptibility just described
through use of the random phase approximation (RPA) of many bod y theory. The Feynman
diagrams included in this method are the same as those incorporated into time dependent
density functional theory, though use of our Hubbard model allow s us to solve the result-
ing equation easily once the very large array of irreducible particle ho le propagators are
generated numerically.
Our task in the present paper is to extend the RPA treatment to inc orporate spin orbit
coupling. The extension is non trivial. The quantity of interest, refe rred to in abbreviated
notation as χ+,−, may be expressed as a commutator of the spin operators S+andS−whose
precise definition is given earlier.5,12With spin orbit coupling ignored, the RPA decoupling
procedure leads to a closed equation for χ+,−. When the RPA decoupling is carried out in
its presence, we are led to a sequence of four coupled equations wh ich include new objects
we may refer to as χ−,−,χ↑,−andχ↓,−. The number of irreducible particle hole propagators
that must be computed likewise is increased by a factor of four. For a very simple version
6of a one band Hubbard model, and for a very different purpose, Fuld e and Luther carried
out an equivalent procedure many years ago15. In what follows, we provide a summary of
key steps along with expressions for the final set of equations.
To generate the equation of motion, we need the commutator of th e operator S+
µν(l,l′) =
c†
lµ↑cl′ν↓with the Hamiltonian. One finds
[S+
µν(l,l′),HSO] =1
2/summationdisplay
η{λl′L+
νηc†
lµ↑cl′η↑−λlL+
ηµc†
lη↓cl′ν↓+λl′Lz
νηc†
lµ↑cl′η↓−λlLz
ηµc†
lη↑cl′ν↓}.(3)
The last two terms on the right hand side of Eq. 3 lead to terms in the e quation of motion
which involve χ+,−whereas the first two terms couple us to the entities χ↑,−andχ↓,−.
When we write down the commutator of these new correlation funct ions with the spin orbit
Hamiltonian, we are led to terms which couple into the function χ−,−which is formed from
the commutator of two S−operators. In the absence of spin orbit coupling, a consequence o f
spin rotation invariance of the Hamiltonian is that the three new func tions just encountered
vanish. But they do not in its presence, and they must be incorpora ted into the analysis.
Onethen introduces theinfluence of theCoulomb interaction into th eequation ofmotion,
and carries out an RPA decoupling of the resulting terms. The analys is is very lengthy, so
here we summarize only the structure that results from this proce dure. Definitions of the
various quantities that enter are given in the Appendix. We express the equations of motion
in terms of a 4 ×4 matrix structure, where in schematic notation we let χ(1)=χ+,−,
χ(2)=χ↑,−,χ(3)=χ↓,−andχ(4)=χ−,−. The four coupled equations then have the form
Ωχ(s)=A(s)+/summationdisplay
s′(Bss′+˜Bss′)χ(s′)(4)
Each quantity in Eq. 4 has attached to it four orbital indices, and fo ur site indices. To be
explicit, χ(2)=χ↑,−which enters Eq. (4) is formed from the commutator of the operat or
c†
lµ↑cl′ν↑withc†
mµ′↓cm′ν′↑and in full we denote this quantity as χ(2)
µν;µ′ν′(ll′;mm′). The site
indices label the planes in the film, and we suppress reference to Ω an d/vectorQ/bardbl. The products
on the right hand side of Eq. 4 are matrix multiplications that involve th ese various indices.
For instance, the object/summationtext
s′Bss′χ(s′)is labeled by four orbital and four site indices so
[Bss′χ(s′)]µν,µ′ν′(ll′;mm′) =/summationdisplay
γδ/summationdisplay
nn′Bss′
µν,γδ(ll′;nn′)χ(s′)
γδ,µ′ν′(nn′;mm′). (5)
One proceeds by writing Eq. 4 in terms of the dynamic susceptibilities t hat characterize
the non-interacting system. These, referred to also as the irred ucible particle hole propaga-
tors, are generated by evaluating the commutators which enter in to the definition of χ(s)in
7the non interacting ground state. These objects, denoted by χ(0s)obey a structure similar
to Eq. 4,
Ωχ(0s)=A(s)+/summationdisplay
s′Bss′χ(s′). (6)
It is then possible to relate χ(s)toχ(0s)through the relation, using four vector notation,
/vector χ(Ω) =/vector χ(0)(Ω)+(Ω −B)−1˜B/vector χ(Ω). (7)
The matrix structure Γ ≡(Ω−B)−1may be generated from the definition of B, which may
be obtained from the equation of motion of the non-interacting sus ceptibility, Eq. 6. Then
˜Bfollows from the equation of motion of the full susceptibility, as gene rated in the RPA.
One may solve Eq. 7
/vector χ(Ω) = [I−(Ω−B)−1¯B]−1/vector χ(0)(Ω), (8)
so our basic task is to compute the non interacting susceptibility mat rix/vector χ(0)and then carry
out the matrix inversion operation displayed in Eq. 8. For this we requ ire the single particle
Greens functions (SPGFs) associated with our approach.
To generate the SPGFs, we set up an effective single particle Hamilton ianHspby intro-
ducing a mean field approximation for the Coulomb interaction. The ge neral structure of
the single particle Hamiltonian is
Hsp=/summationdisplay
ij/summationdisplay
µνσ˜Tµνσ
ijc†
iµσcjνσ+/summationdisplay
i/summationdisplay
µν{α∗
i;µνc†
iµ↓ciν↑+αi;µνc†
iµ↑ciν↓} (9)
where the effective hopping integral ˜Tµνσ
ijcontains the spin diagonal portion of the spin orbit
interaction, along withthemeanfield contributions fromtheCoulomb interaction. The form
we use for the latter is stated below. The coefficients in the spin flip te rms are given by
αi;µν=λiL−
µν−/summationdisplay
ηγUi;ηµ,νγ/an}bracketle{tc†
iη↓ciγ↑/an}bracketri}ht. (10)
We then have the eigenvalue equation that generates the single par ticle eigenvalues and
eigenfunctions in the form Hsp|φs/an}bracketri}ht=Es|φs/an}bracketri}ht; we can write this in the explicit form
/summationdisplay
l/summationdisplay
ησ′/bracketleftBig
δσσ′˜Tµησ′
il+δil(δσ′↓δσ↑α∗
l;µη+δσ′↑δσ↓αl;µη)/bracketrightBig
/an}bracketle{tlησ′|φs/an}bracketri}ht=Es/an}bracketle{tiµσ|φs/an}bracketri}ht.(11)
The single particle Greens function may be expressed in terms of the quantities that enter
Eq. 11. We have for this object the definition
Giµσ;jνσ′(t) =−iθ(t)/an}bracketle{t{ciµσ(t),c†
jνσ′(0)}/an}bracketri}ht (12)
8and one has the representation
Giµσ;jνσ′(Ω) =/summationdisplay
s/an}bracketle{tiµσ|φs/an}bracketri}ht/an}bracketle{tφs|jνσ′/an}bracketri}ht
Ω−Es+iη. (13)
These functions may be constructed directly from their equations of motion, which read,
after Fourier transforming with respect to time,
−/summationdisplay
l/summationdisplay
ησ′′/bracketleftBig
δσσ′′˜Tµησ′′
il+δil(δσ′′↓δσ↑α∗
l;µη+δσ′′↑δσ↓αl;µη)/bracketrightBig
Glησ′′;jνσ′+ΩGiµσ;jνσ′=δσσ′δµνδij.
(14)
Forthecasewherethesubstrateissemi infinite, ourmeansofgen eratinganumerical solution
to the hierarchy of equations stated in Eq. 14 has been discussed e arlier. What remains is
to describe how the Coulomb interaction enters the effective hoppin g integrals ˜Tµνσ
ijthat
appear in Eq. 9, Eq. 11 and Eq. 14.
There are, of course, a large number of Coulomb matrix elements in t he original Hamil-
tonian, even if the Coulomb interactions are confined to within the 3d shell. Through the
use of group theory,17the complete set of Coulomb matrix elements may be expressed in
terms of three parameters. These are given in Table I of the first c ited paper in ref. 5.
In subsequent work, we have found that a much simpler structure18nicely reproduces re-
sults obtained with the full three parameter form. We use the simple r one parameter form
here, for which Ui;µν,µ′ν′=Uiδµν′δµ′ν. Then in the mean field approximation, the Coulomb
contribution to the single particle Hamiltonian assumes the form
H(C)
sp=−/summationdisplay
iUimi
2/summationdisplay
µ(c†
iµ↑ciµ↑−c†
iµ↓ciµ↓) (15)
Heremiis the magnitude of the moment on site i. The Coulomb interactions Uiare non zero
only within the ultrathin ferromagnet, and the moments mi, determined self consistently,
vary from layer to layer when we consider multi layer ferromagnetic films.
It should be noted that when the Ansatz just described is employed in Eq. 10, the term
from the Coulomb interaction on the right hand side becomes propor tional to the transverse
component of the moment located on site iand this vanishes identically. Thus, despite the
complexity introduced by the spin orbit coupling, when the simple one p arameter Ansatz for
the Coulomb matrix elements is employed, one needs no parameters b eyond the moment on
each layer in the self consistent loops that describe the ground sta te. In the present context,
this is an extraordinarily large savings in computational labor, and th is will allow us to
9address very large systems in the future. It is the case that cert ain off diagonal elements
such as/an}bracketle{tc†
mµ′↓cl′ν↑/an}bracketri}htappear in the quantities defined in the Appendix. Notice, for example ,
the expressions in Eqs. A.1. Once the ground state single particle Gr eens functions are
determined, such expectation values are readily computed.
III. RESULTS AND DISCUSSION
In earlier studies of Fe layers on W(110),19,20as noted above, the electronic structure
was generated through use of tight binding parameters obtained f rom bulk electronic struc-
ture calculations. These calculations generate effective exchange interactions comparable
in magnitude to those found in the bulk transition metals,20with the consequence that for
both monolayer Fe and bilayer Fe on W(110) the large wave vector sp in waves generated
by theory are very much stiffer than found experimentally21,22though it should be noted
that for the bilayer, the calculated value of the spin wave exchange stiffness is in excellent
accord with the data.23Subsequent calculations which construct the spin wave dispersion
relation from adiabatic theory based on calculations of effective exc hange integrals also gen-
erate spin waves for the monolayer substantially stiffer than found experimentally,12though
they are softer than in our earlier work by a factor of two or so. We remark that it has
been suggested that the remarkably soft spin waves found exper imentally may have origin
in carbon contamination of the monolayer and bilayer.20We remark here that this can be
introduced during the SPEELS measurement. We note that the mag netic properties of Fe
monolayers grown on carbon free W(110)24differ dramatically from those grown on surfaces
now known to be contaminated by carbon.25In the former case, the domain walls have a
thickness of 2.15 nm,24whereas in the latter circumstance very narrow walls with thickness
bounded from above by 0.6 nm are found.25This suggests that the strength of the effective
exchange is very different in the two cases, with stiffer exchange in t he carbon free samples.
The considerations of the previous paragraph have motivated us t o carry out a series of
studies of the effective exchange in the Fe monolayer on W(110) with in the framework of
three different electronic structure calculations. We find that alth ough all three give local
density of states that are very similar, along with very similar energy bands when these are
examined, theintersiteexchangeinteractionsvarysubstantially. First, wehaveemployed the
parameter set used earlier that is based on bulk electronic structu res19,20in new calculations
10we call case A. In case B, we have employed an approach very similar t o that used in ref. 12,
though in what follows our calculation of effective exchange integrals is non relativistic.
This is the Korringa Kohn Rostoker Greens Function (KKR-GF) meth od,26which employs
the atomic sphere approximation and makes use of the Dyson equat ionG=g+gVGas
given in matrix notation. This allows us to calculate the Greens functio nGof an arbitrary
complex system given the perturbing potential Vand the Greens function gof a reference
unperturbed system. Within the Local Spin Density Approximation ( LSDA),27We consider
a slab of five monolayers of W with the experimental lattice constant on top of which an Fe
monolayer is deposited and relaxed by -12.9%12with respect to the W interlayer distance.
Angular momenta up to lmax= 3 were included in the Greens functions with a kmesh of
6400 points in the full two dimensional Brillouin zone. The effective exc hange interactions
were calculated within the approximation of infinitesmal rotations28that allows one to use
the magnetic force theorem. This states that the energy change due to infinitesmal rotations
in the moment directions can be calculated through the Kohn Sham eig envalues.
Method C is the Real Space Linear-Muffin-Tin-Orbital approach as im plemented, also,
in the atomic sphere approximation (RS-LMTO-ASA).29–33Due to its linear scaling, this
method allows one to address the electronic structure of systems with a large number of
atoms for which the basic eigenvalue problem is solved in real space us ing the Haydock
recursion method. The Fe overlayer on the W(110) substrate was simulated by a large
bcc slab which contained ∼6800 atoms, arranged in 12 atomic planes parallel to the (110)
surface, with the experimental lattice parameter of bulk W. One em pty sphere overlayer
is included, and self consistent potential parameters were obtaine d for the empty sphere
overlayer, the Fe monolayer, and the three W layers underneath u sing LSDA.34For deeper
W layers we use bulk potential. Nine orbitals per site (the five 3d and 4 s p complex) were
used to describe the Fe valence band and the empty sphere overlay er, and for W the fully
occupied 4f orbitals were also included in the core. To evaluate the or bital moments we use
a scalar relativistic (SR) approach and include a spin orbit coupling ter mλ/vectorL·/vectorSat each
variational step.35In the recursion method the continued fraction has been terminat ed after
30 recursion levels with the Beer Pettifor terminator.36The TB parameters so obtained are
inserted into our semi-empirical scheme and this allows us to generat e the non interacting
susceptibilities which enter our full RPA description of the response of the structure.
In order to compare the electronic structures generated by the approaches just described,
11we turn our attention to the local density of states for the major ity and minority spins in
the adsorbed Fe monolayer. These are summarized in Fig. 1.
The local densities of states (LDOS) generated by the three sets of TB parameters have
approximately the same overall features, as we see from Fig. 1. Th e main differences appear
in the majority spin band, which overlaps the 5d states in the W subst rate over a larger
energy range than the minority band. This is also true if we compare t he LDOS generated
by the tight binding parameters extracted from the KKR electronic structure to the LDOS
obtaineddirectly fromtheKKR calculations (reddashed linein Fig.1b) . The Fe-Whopping
parameters are indeed the least accurate portion of our paramet rization scheme. In case A
wejustusedtheFe-FebulkparameterstodescribetheFe-Whopp ing. IncaseBweextracted
TB parameters for Fe by fitting a KKR calculation of an unsupported Fe monolayer with a
lattice parameter matching that of the W substrate. For the Fe-W hopping we used the Fe
parameters obtained from the fitting, scaled to mimic the Fe-W dista nce relaxation. The
relaxation parameter was chosen to give the correct spin magnetic moment for the adsorbed
Fe monolayer. In case C all parameters were directly provided by th e RS-LMTO-ASA code,
but in the DFT calculations the Fe-W distance was assumed to be equa l to the distance
between W layers. Thus, the main difference between cases B and C is the treatment of the
mixing between Fe and W states and this is expected to affect more st rongly the states that
occupy the same energy range.
As noted above, while the local density of states provided by the th ree approaches to
the electronic structure are quite similar as we see from Fig. 1 (and t he same is true of the
electronic energy bands themselves if these are examined), the ex change interactions differ
substantially for the three cases. For the first, second and third neighbors we have (in meV)
42.5, 3.72 and 0.46 for model A, 28.7, -7.87 and 0.31, for model B and 1 1.23, -7.31 and 0.22
for model C. The authors of ref. 12 find 10.84, -3.34 and 3.64 for th ese exchange integrals.
We now turn to our studies of spin excitations in the Fe monolayer and Fe bilayer on
W(110) within the framework of the electronic structure generat ed through use of the ap-
proach in case C. We will discuss the influence of spin orbit coupling on b oth the trans-
verse wave vector dependent susceptibility though study of the s pectral density function
A(/vectorQ/bardbl,Ω;l⊥) =−Im{χ+,−(/vectorQ/bardbl,Ω;l⊥)}discussed in section I. This function, for fixed wave
vector/vectorQ/bardbl, when considered as a function of frequency Ω, describes the fre quency spectrum
of the fluctuations of wave vector /vectorQ/bardblof the transverse magnetic moment in layer l⊥as noted
12-6 -4 -2 0 2
E-EF (eV)-2024LDOS (states/eV)-2024LDOS (states/eV)-2024LDOS (states/eV)a)
b)
c)
FIG. 1: (color online) For the Fe monolayer adsorbed on W(110 ), we show the local density of
states in the Fe monolayer. The majority spin density of stat es is shown positive and the minority
spin density of states is negative. The zero of energy is at th e Fermi energy. In (a), bulk electronic
structure parameters are used as in the second of the two pape rs cited in 5 (caseA). In (b), we
have the density of states generated by method B. The black cu rve is found by fitting the KKR
electronic energy band structure to tight binding paramete rs as described in the text, and the red
curve is calculated directly from the KKR calculation. In (c ) we have the local density of states
generated by method C.
13above. In the frequency regime where spin waves are encountere d, this function is closely
related to (but not identical to) the response function probed in a SPEELS measurement.
In Fig. 2, for the Fe monolayer on W(110), we show the spectral de nsity function cal-
culated for three values of |/vectorQ/bardbl|, for propagation across the magnetization. Thus, the wave
vector is directed along the short axis in the surface. This is the dire ction probedin SPEELS
studies ofthe Femonolayer onthissurface.22Ineach figure, we show three curves. The green
dashed curve is calculated with spin orbit coupling set to zero. We sho w only a single curve
for this case, because the spectral density is identical for the tw o directions of propagation
across the magnetization, + /vectorQ/bardbland−/vectorQ/bardbl. When spin orbit coupling is switched on, for the
two directions just mentioned the response function is very differe nt, as we see from the red
andblack curve inthevariouspanels. These spin wave frequencies, deduced fromthepeakin
the response functions as discussed in section I, differ for the two directions of propagation,
and also note that the peak intensities and linewidths differ as well. It is the absence of both
time reversal symmetry and reflection symmetry which renders + /vectorQ/bardbland−/vectorQ/bardblinequivalent
for this direction of propagation. The system senses this breakdo wn of symmetry through
the spin orbit interaction. If one considers propagation parallel to the magnetization, the
asymmetries displayed in Fig. 2 are absent. The reason is that for th is direction of propaga-
tion, reflection in the plane that is perpendicular to both the magnet ization and the surface
is a goodsymmetry operation of the system, but takes + /vectorQ/bardblinto−/vectorQ/bardblthus rendering the two
directions equivalent. Recall, of course, that the magnetization is a pseudo vector in regard
to reflections. Notice how very broad the curves are for large wav e vectors; the lifetime of
the spin waves is very short indeed.
As discussed in section I, we may construct a spin wave dispersion cu rve by plotting the
maxima in spectral density plots such as those illustrated in Fig. 2 as a function of wave
vector. We show dispersion relations constructed in this manner in F ig. 3, with spin orbit
coupling both present and absent. In Fig. 3a, and for propagation perpendicular to the
magnetization we show the dispersion curve so obtained for wave ve ctors throughout the
surface Brillouin zone, and in Fig. 3b we show its behavior for small wav e vectors.
Let is first consider Fig, 3(a). Here the dispersion curve extends t hroughout the two
dimensional Brillouin zone. At the zone boundary, quite clearly the slo pe of the dispersion
curve does not vanish. In this direction of propagation, the natur e of the point at the zone
boundary does not require the slope to vanish. What is most striking , clearly, is the anomaly
140 10 20 30 40 50
Ω (meV)05×1031×104-Imχ+,−(a)
0 50 100 150 200
Ω (meV)050010001500-Imχ+,-(b)
0 100 200 300
Ω (meV)0200400600800-Imχ+,-(c)
FIG. 2: (color online) Thespectral density functions A(/vectorQ/bardbl,Ω;l⊥) evaluated in theFe monolayer for
three values of the wave vector in the direction perpendicul ar to the magnetization. We have (a)
|/vectorQ/bardbl|= 0.4˚A−1, (b)|/vectorQ/bardbl|= 1.0˚A−1and (c)|/vectorQ/bardbl|= 1.4˚A−1. The green curve (dashed) is calculated
with spin orbit coupling set to zero; the spectral density he re is independent of the sign of /vectorQ/bardbl. The
red and black curves are calculated with spin orbit coupling turned on. Now we see asymmetries
for propagation across the magnetization, with the red curv e/vectorQ/bardbldirected from left to right and the
black curve from right to left.
in the vicinity of 1 ˚A−1. This feature is evident in the calculation with spin orbit coupling
absent, and for positive values of the wave vector the feature be comes much more dramatic
when spin orbit coupling is switched on. Anomalies rather similar to thos e in the black
curve in Fig. 3(a) appear in the green dispersion curve found in Fig. 3 of ref. 12, though
these authors did not continue their calculation much beyond the 1 ˚A−1regime. Our spin
waves are very much softer than theirs in this spectral region, no tice. In Fig. 3 of ref. 12,
one finds two dispersion curves, one a mirror image of the second. T hus, these authors
15-1.5 -1-0.5 00.5 11.5
Q (A-1)050100150200Ω (meV)(a)
-0.4 -0.2 0 0.2 0.4
Q (A-1)010203040Ω (meV)(b)
FIG. 3: (color online) Spin wave dispersion relations const ructed from peaks in the spectral den-
sity, for the Fe monolayer on W(110). The wave vector is in the direction perpendicular to the
magnetization. The red curve is constructed in the absence o f spin orbit coupling, it is included in
the black curve.
display two spin wave frequencies for each wave vector. This surely is not correct. For a
structure with one atom per unit cell, there is one and only one spin wa ve mode for each
wave vector, though as discussed above for the structure explo red here symmetry allow the
left/right asymmetry in the dispersion curve illustrated in our Fig. 3.
In Fig. 3b, again with spin orbit coupling switched on and off, we show an expanded
view of the dispersion curve for small wave vectors. With spin orbit in teraction switched off,
at zero wave vector we see a zero frequency spin wave mode, as re quired by the Goldstone
theorem when the underlying Hamiltonian is form invariant under spin r otation. The curve
is also symmetrical, and is accurately fitted by the form Ω( /vectorQ/bardbl) = 149Q2
/bardbl(meV), with the
wave vector in ˚A−1, whereas with spin orbit coupling turned on the dispersion relation is
fitted by Ω( /vectorQ/bardbl) = 3.4−11.8Q/bardbl+143Q2
/bardbl(meV). Spin orbit coupling introduces an anisotropy
gap atQ/bardbl= 0, and most striking is the term linear in wave vector. This has its orig in in the
Dzyaloshinskii Moriya interaction whose presence, as argued by th e authors of ref. 9, has its
origin in the absence of both time reversal and inversion symmetry, for the adsorbed layer.
At long wavelengths, one may describe spin waves by classical long wa velength phe-
nomenology. The linear term in the dispersion curve has its origin in a te rm in the energy
density of the spin system of the form
VDM=−Γ/integraldisplay
dxdzS y(x,z)∂Sx(x,z)
∂x(16)
16HereSα(x,z) is a spin density, the xzplane is parallel to the surface, and the magnetization
is parallel to the zdirection.
One interesting feature of the spin wave mode whose dispersion rela tion is illustrated in
Fig. 3b is that at Q/bardbl= 0, the mode has a finite group velocity. The fit to the dispersion
curve gives this group velocity to be∂Ω(/vectorQ/bardbl)
∂Q/bardbl≈2×105cm/s , which is in the range of acoustic
phonon group velocities.
Weturnnowtoourcalculationsofspinwavesandtheresponsefunc tionsfortheFebilayer
on W(110). Let us first note that experimentally the orientation of the magnetization in
the bilayer appears to be dependent on the surface upon which the bilayer is grown. For
instance, when the bilayer is on the stepped W(110) surface, it is ma gnetized perpendicular
to the surface,25a result inagreement with abinitio calculations of the anisotropy realiz ed in
the epitaxial bilayer.38However, in the SPEELS studies of spin excitations in the bilayer22,39
the magnetization is in plane. In our calculations, we find for model B t he magnetization is
perpendicular to the surface, whereas in model C it lies in plane, along the long axis very
much as in the SPEELS experiments. The anisotropy in the bilayer is no t particularly large,
on the order of 0.5 meV/Fe atom, and one sees from these results t hat it is a property quite
sensitive to the details of the electronic structure. The fact that model B and model C give
the two different stable orientation of the magnetization allows us to explore spin excitations
for the two different orientations of the magnetization.
We first turn our attention to the case where the magnetization lies in plane. The bilayer
has two spin wave modes, an acoustic mode for which the magnetizat ion in the two planes
precesses in phase, and an optical mode for which they precess 18 0 degrees out of phase.
In Fig. 4, we show calculations of the dynamic susceptibility in the freq uency range of the
acoustic mode for two values of the wave vector, Q/bardbl= +0.5˚A−1 andQ/bardbl=−0.5˚A−1. A
spin orbit induced left right asymmetry is clearly evident both in the pe ak frequency and
the height of the feature. Very recently, beautiful measuremen ts of spin orbit asymmetries
in the Fe bilayer have appeared,39and the results of our Fig. 4 are to be compared with
Fig. 3 of ref. 39. Theory and experiment are very similar, both in reg ard to the intensity
asymmetry and also the spin orbit induced frequency shift, though our calculated spin wave
frequencies are a little stiffer than those found experimentally.
As remarked above, in Fig. 4 we show only the acoustical spin wave mo de frequency
regime. In Fig. 5, for the spectral densities in the innermost layer ( upper panel) and the
170 50 100
Ω (meV)050010001500200025003000-Imχ+,−
FIG. 4: (color online) The spectral density in the innermost layer, in the acoustic spin wave regime,
for wave vectors of Q/bardbl= +0.5˚A−1(black curve) and Q/bardbl=−0.5˚A−1(red curve). Model C has
been used for the calculation. In the ground state, the magne tization lies in plane along the long
axis.
outermost layer (lower panel) we show the spectral densities for t he entire spin wave regime,
including the region where the optical spin wave is found. It is clear th at the spin orbit
inducedfrequency shiftsarelargestfortheopticalmodewhich, u nfortunatelyisnotobserved
in the experiments.39
In Fig. 6 for a sequence of wave vectors, all chosen positive, we sh ow a sequence of spectra
calculated for the entire frequency range so both the acoustic an d optical spin wave feature
are displayed. The black curves show the spectral density of the in nermost Fe layer, and
the red curves are for the outer layer. The optical spin wave mode , not evident in the data,
shows clearly in these figures. Notice that for wave vectors great er than 1 ˚A−1the acoustical
mode is localized in the outer layer and the optical mode is localized on th e inner layer.
The optical mode is very much broader than the acoustical mode at large wave vectors, by
virtue of the strong coupling to the electron hole pairs in the W 5d ban ds.
Aninteresting issue is the absence of the optical mode fromthe SPE ELS spectra reported
18-1500
-1000
-500
0-Imχ+,−
0 50 100 150 200 250
Ω (meV)-2000
-1000
0-Imχ+,−
FIG. 5: (color online) For the wave vector Q/bardbl= 0.5˚A−1we show the spectral densities in the
innermost Fe layer (upper panel) and in the outermost layer ( lower panel) for the Fe bilayer on
W(110). Thefigureincludestheoptical spinwave feature. As inFig. 4, theblack curveiscalculated
forQ/bardblpositive, and the red curves are for Q/bardblnegative. The calculations employ model C.
in refs. 22 and 39. We note that these spectra are taken with only t wo beam energies, 4 eV
and 6.75 eV. At such very low energies, the beam electron will sample b oth Fe layers,
so the SPEELS signal will be a coherent superposition of electron wa ves backscattered
from each layer; the excitation process involves coherent excitat ion of both layers by the
incident electron. As a consequence of the 180 degree phase differ ence in spin motions
associated with the two modes it is quite possible, indeed even probab le, that for energies
where the acoustical mode is strong the intensity of the optical mo de is weak, by virtue
of quantum interference effects in the excitation scattering amplit ude. In earlier studies
of surface phonons, it is well documented that on surfaces where two surface phonons of
different polarization exist for the same wave vector, one can be sile nt and one active in
electron loss spectroscopy.40It would require a full multiple scattering analysis of the spin
waveexcitationprocesstoexplorethistheoretically. Whileearlier41calculationsthataddress
SPEELS excitation of spin waves described by the Heisenberg model could be adapted for
190200040006000Imχ+,−
010002000Imχ+,−
05001000Imχ+,−
050010001500Imχ+,−
050010001500Imχ+,−
0 100 200 300 400 500
Ω (meV)05001000Imχ+,−Qy=0.4 A-1
Qy=0.6 A-1
Qy=0.8 A-1
Qy=1.0 A-1
Qy=1.2 A-1
Qy=1.4 A-1
FIG. 6: (color online) For the Fe bilayer and for several valu es of the wave vector (all positive), we
show the spectral density functions for the innermost layer adjacent to the substrate (black curve)
and those for the outer layer of the film. The calculations emp loy model C.
20-1.5 -1-0.5 00.5 11.5
Q (A-1)0100200300Ω (meV)
FIG. 7: (color online) For the Fe bilayer with magnetization in plane, we show the spin wave
dispersion curves calculated with spin orbit coupling (bla ck points) and without spin orbit coupling
(red curves). Model C has been employed for these calculatio ns.
this purpose, in principle, a problem is that at such low beam energies it is necessary to take
due account of image potential effects to obtain meaningful result s.42This is very difficult
to do without considerable information on the electron reflectivity o f the surface.42It would
be of great interest to see experimental SPEELS studies of the Fe bilayer with a wider range
of beam energies to search for the optical mode, if this were possib le.
In Fig. 7, we show dispersion curves for the optical and acoustic sp in wave branches for
the bilayer. The magnetization lies in plane, and one can see that on th e scale of this figure,
the spin orbit effects on the dispersion curve are rather modest co mpared to those in the
monolayer. For small wave vectors, with spin orbit coupling present , the dispersion curve of
the acoustic spin wave branch is fitted by the form Ω( Q/bardbl) = 0.49−0.85Q/bardbl+243Q2
/bardbl(meV)
so at long wavelengths the influence of the Dzyaloshinskii Moriya inte raction is more than
one order of magnitude smaller than it is in the monolayer.
If the magnetization is perpendicular to the surface, then symmet ry considerations show
that there are no left/right asymmetries in the spin wave propagat ion characteristics. One
may see this as follows. Consider a wave vector /vectorQ/bardblin the plane of the surface, which
also is perpendicular to the magnetization, and thus perpendicular t o the long axis. The
reflection Rin the plane perpendicular to the surface and which contains the mag netization
210 100 200 300 400
Ω (meV)050010001500200025003000Imχ+,−
FIG. 8: (color online) For the bilayer and the case where the m agnetization is perpendicular to the
surface (model B), and for Q/bardbl= 0.6˚A−1, we show spectral density function calculated for positive
values of Q/bardbl(continuous lines) and negative values of Q/bardbl(symbols). The black curve is the spectral
density for the inner layer, and the red curve is the outermos t Fe layer.
simultaneously changes the sign of wave vector and the magnetizat ion. If this is followed
by the time reversal operation T, then/vectorQ/bardblremains reversed in sign but the magnetization
changes back to its original orientation. Thus the product RTleaves the system invariant
but transforms /vectorQ/bardblinto−/vectorQ/bardbl. The two propagation directions are then equivalent.
We illustrate this in Fig. 8 where, for Q/bardbl= 0.6˚A−1, where it is shown that the spectral
densities calculated for the two directions of propagation are ident ical, with spin orbit cou-
pling switched on. Model B, in which the magnetization is perpendicular to the surface, has
been used in these calculations. The spectral densities calculated f or the two signs of Q/bardbl
cannot be distinguished to within the numerical precision we use.
IV. CONCLUDING REMARKS
We have developed the formalism which allows one to include the influenc e of spin orbit
couplingonthespinexcitationsofultrathinferromagnetsonsemiin finitemetallicsubstrates.
Our approach allows us to calculate the full dynamic susceptibility of t he system, so as
illustrated by the calculations presented in section III we can examin e the influence of spin
orbit coupling on the linewidth (or lifetime) of spin excitations, along wit h their oscillator
22strength. As in previous work, we can then construct effective dis persion curves by following
peaks in the spectral density as a function of wave vector, withou t resort to calculations of
large numbers of very small distant neighbor exchange interaction s. The results presented
in Fig. 4 are very similar to the experimental data reported in ref. 39 , as discussed above,
though we see that in the bilayer the influence of the Dzyaloshinskii M oriya interaction is
considerably more modest than in the monolayer.
We will be exploring other issues in the near future. One interest in ou r minds is the
influence of spin orbit coupling on the spin pumping contributions to th e ferromagnetic
resonance linewidth, as observed in ferromagnetic resonance (FM R) studies of ultrathin
films.43It has been shown earlier44that the methodology employed in the present paper
(without spin orbit coupling included) can be applied to the description of the spin pumping
contribution to the FMR linewidth, and in fact an excellent quantitativ e account of the
data on the Fe/Au(100) system was obtained. It is possible that fo r films grown on 4d
and 5d substrates that spin orbit coupling can influence the spin pum ping relaxation rate
substantially. This willrequirecalculationsdirected towardmuchthic ker filmsthanexplored
here. The formalism we have developed and described in the present paper will allow such
studies in the future.
Acknowledgments
This research was supported by the U. S. Department of Energy, through grant No.
DE-FG03-84ER-45083. S. L. wishes to thank the Alexander von Hu mboldt Foundation
for a Feodor Lynen Fellowship. A.T.C. and R.B.M. acknowledge support from CNPq and
FAPERJ and A.B.K. was supported also by the CNPq, Brazil.
Appendix
In this Appendix we provide explicit expressions for the various quan tities which enter
the equations displayed in Section II. While these expressions are un fortunately lengthy, it
will be useful for them to be given in full.
A(1)
µν,µ′ν′(ll′;mm′) =δl′mδνµ′/an}bracketle{tc†
lµ↑cm′ν′↑/an}bracketri}ht−δlm′δµν′/an}bracketle{tc†
mµ′↓cl′ν↓/an}bracketri}ht
23A(2)
µν,µ′ν′(ll′;mm′) =−δlm′δµν′/an}bracketle{tc†
mµ↓cl′ν↑/an}bracketri}ht
A(3)
µν,µ′ν′(ll′;mm′) =δl′mδνµ′/an}bracketle{tc†
lµ↓cm′ν′↑/an}bracketri}ht
A(4)
µν,µ′ν′(ll′;mm′) = 0 (A.1)
The various expectation values in the equations above and those dis played below are
calculated from the single particle Greens functions once the self co nsistent ground state
parameters are determined. Then
˜B11
µν,µ′ν′(ll′;mm′) =/summationdisplay
η/parenleftBig
Ul;µ′η,µν′/an}bracketle{tc†
lη↓cl′ν↓/an}bracketri}htδlmδlm′−Ul′;µ′ν,ην′/an}bracketle{tc†
lµ↑cl′η↑/an}bracketri}htδl′mδl′m′/parenrightBig
˜B12
µν,µ′ν′(ll′;mm′) =/summationdisplay
η/parenleftBig
Ul′;µ′ν,ν′η/an}bracketle{tc†
lµ↑cl′η↓/an}bracketri}htδl′mδl′m′−Ul;ηµ′,µν′/an}bracketle{tc†
lη↑cl′ν↓/an}bracketri}htδlmδlm′+
+Ul;µ′η,µν′/an}bracketle{tc†
lη↑cl′ν↓/an}bracketri}htδlmδlm′/parenrightBig
˜B13
µν,µ′ν′(ll′;mm′) =/summationdisplay
η/parenleftBig
Ul′;µ′ν,ν′η/an}bracketle{tc†
lµ↑cl′η↓/an}bracketri}htδl′mδl′m′−Ul;ηµ′,µν′/an}bracketle{tc†
lη↑cl′ν↓/an}bracketri}htδlmδlm′−
−Ul′;µ′ν,ην′/an}bracketle{tc†
lµ↑cl′η↓/an}bracketri}htδl′mδl′m′/parenrightBig
˜B14
µν,µ′ν′(ll′;mm′) = 0
(A.2)
˜B21
µν,µ′ν′(ll′;mm′) =/summationdisplay
ηUl;µ′η,µν′/an}bracketle{tc†
lη↓cl′ν↑/an}bracketri}htδlmδlm′
˜B22
µν,µ′ν′(ll′;mm′) =/summationdisplay
η/bracketleftBig
(Ul′;µ′ν,ν′η−Ul′;µ′ν,ην′)/an}bracketle{tc†
lµ↑cl′η↑/an}bracketri}htδl′mδl′m′−
−(Ul;ηµ′,µν′−Ul;µ′η,µν′)/an}bracketle{tc†
lη↑cl′ν↑/an}bracketri}htδlmδlm′/bracketrightBig
˜B23
µν,µ′ν′(ll′;mm′) =/summationdisplay
η/parenleftBig
Ul;µ′ν,ν′η/an}bracketle{tc†
lµ↑cl′η↑/an}bracketri}htδl′mδl′m′−Ul;ηµ′,µν′/an}bracketle{tc†
lη↑cl′ν↑/an}bracketri}htδlmδl′m′/parenrightBig
˜B24
µν,µ′ν′(ll′;mm′) =−/summationdisplay
ηUl′;µ′ν,ην′/an}bracketle{tc†
lµ↑cl′η↓/an}bracketri}htδl′mδl′m′
(A.3)
˜B31
µν,µ′ν′(ll′;mm′) =−/summationdisplay
ηUl′;µ′ν,ην′/an}bracketle{tc†
lµ↓cl′η↑/an}bracketri}htδl′mδl′m′
˜B32
µν,µ′ν′(ll′;mm′) =/summationdisplay
η/parenleftBig
Ul;µ′ν,ν′η/an}bracketle{tc†
lµ↓cl′η↓/an}bracketri}htδl′mδl′m′−Ul;ηµ′,µν′/an}bracketle{tc†
lη↓cl′ν↓/an}bracketri}htδlmδl′m′/parenrightBig
24˜B33
µν,µ′ν′(ll′;mm′) =/summationdisplay
η/bracketleftBig
(Ul′;µ′ν,ν′η−Ul′;µ′ν,ην′)/an}bracketle{tc†
lµ↓cl′η↓/an}bracketri}htδl′mδl′m′−
−(Ul;ηµ′,µν′−Ul;µ′η,µν′)/an}bracketle{tc†
lη↓cl′ν↓/an}bracketri}htδlmδlm′/bracketrightBig
˜B34
µν,µ′ν′(ll′;mm′) =/summationdisplay
ηUl;µ′η,µν′/an}bracketle{tc†
lη↑cl′ν↓/an}bracketri}htδlmδlm′
(A.4)
˜B41
µν,µ′ν′(ll′;mm′) = 0
˜B42
µν,µ′ν′(ll′;mm′) =/summationdisplay
η/parenleftBig
Ul′;µ′ν,ν′η/an}bracketle{tc†
lµ↓cl′η↑/an}bracketri}htδl′mδl′m′−Ul;ηµ′,µν′/an}bracketle{tc†
lη↓cl′ν↑/an}bracketri}htδlmδlm′−
−Ul′;µ′ν,ην′/an}bracketle{tc†
lµ↓cl′η↑/an}bracketri}htδl′mδl′m′/parenrightBig
˜B43
µν,µ′ν′(ll′;mm′) =/summationdisplay
η/parenleftBig
Ul′;µ′ν,ν′η/an}bracketle{tc†
lµ↓cl′η↑/an}bracketri}htδl′mδl′m′−Ul;ηµ′,µν′/an}bracketle{tc†
lη↓cl′ν↑/an}bracketri}htδlmδlm′+
+Ul;µ′η,µν′/an}bracketle{tc†
lη↓cl′ν↑/an}bracketri}htδlmδlm′/parenrightBig
˜B44
µν,µ′ν′(ll′;mm′) =/summationdisplay
η/parenleftBig
Ul;µ′η,µν′/an}bracketle{tc†
lη↑cl′ν↑/an}bracketri}htδlmδl′m−Ul′;µ′ν,ην′/an}bracketle{tc†
lµ↓cl′η↓/an}bracketri}htδl′mδl′m′/parenrightBig
(A.5)
B11
µν,µ′ν′(ll′;mm′) =˜Tνν′↓
l′m′δlmδµµ′−(˜Tµµ′↑
lm)∗δl′m′δνν′
B12
µν,µ′ν′(ll′;mm′) =α∗
l′;ν′νδlmδl′m′δµµ′
B13
µν,µ′ν′(ll′;mm′) =−α∗
l;µµ′δlmδl′m′δνν′
B14
µν,µ′ν′(ll′;mm′) = 0 (A.6)
B21
µν,µ′ν′(ll′;mm′) =αl′;νν′δlmδl′m′δµµ′
B22
µν,µ′ν′(ll′;mm′) =˜Tνν′↑
l′m′δlmδµµ′−(˜Tµµ′↑
lm)∗δl′m′δνν′
B23
µν,µ′ν′(ll′;mm′) = 0
B24
µν,µ′ν′(ll′;mm′) =−α∗
l;µµ′δlmδl′m′δνν′ (A.7)
B31
µν,µ′ν′(ll′;mm′) =−αl;µ′µδlmδl′m′δνν′
25B32
µν,µ′ν′(ll′;mm′) = 0
B33
µν,µ′ν′(ll′;mm′) =˜Tνν′↓
l′m′δlmδµµ′−(˜Tµµ′↓
lm)∗δl′m′δνν′
B34
µν,µ′ν′(ll′;mm′) =α∗
l′;ν′νδlmδl′m′δµµ′ (A.8)
B41
µν,µ′ν′(ll′;mm′) = 0
B42
µν,µ′ν′(ll′;mm′) =−αl;µ′µδlmδl′m′δνν′
B43
µν,µ′ν′(ll′;mm′) =αl′;νν′δlmδl′m′δµµ′
B11
µν,µ′ν′(ll′;mm′) =˜Tνν′↑
l′m′δlmδµµ′−(˜Tµµ′↓
lm)∗δl′m′δνν′ (A.9)
1Rodrigo Arias and D. L. Mills, Phys. Rev. B 60, 7395 (1999); A. Azevedo, A. B. Oliveira, F. M.
de Aguiar, and S. M. Rezende, Phys. Rev. B 62, 5331 (2000); D. L. Mills and S. M. Rezende
in Spin Dynamics in Confined Structures II, (Springer Verlag , Heidelberg, 2002); P. Landeros,
Rodrigo E. Arias and D. L. Mills, Phys. Rev. B 77, 214405, (2008); J. Lindner, I. Barsukov, C.
Raeder, C. Hassel, O. Posth, R. Meckenstock, P. Landeros and D. L. Mills, Phys. Rev. B 80,
224421 (2009).
2R.Urban, G. Woltersdorf andG. Heinrich, Phys.Rev. Lett. 87, 217204 (2001), Y.Tserkovnyak,
A. Brataas and G. E. W. Bauer, Phys. Rev. Lett. 88, 117601 (2002); Phys. Rev. B 66, 224403
(2002); D. L. Mills, Phys. Rev. B 68, 014419 (2003); M. Zwierzycki, Y. Tserkovnyak, P. J.
Kelly, A. Brataas, and G. E. W. Bauer, Phys. Rev. B 71, 064420 (2005); A. T. Costa, R. B.
Muniz and D. L. Mills, Phys. Rev. B 73, 054426 (2006).
3S. M. Rezende, A. Azevedo, M. A. Lucena and F. M. Aguiar, Phys. Rev. B63, 214416 (2001).
4R. Vollmer, M. Etzkorn, P. S. Anil Kumar, H. Ibach and J. R. Kir schner, Phys. Rev. Lett.
91, 14720 (2003); for a theoretical discussion of the linewidt h and dispersion observed in the
experiment, see A. T. Costa, R. B. Muniz Phys. Rev. B 70, 0544406 (2004).
5Early papers are H. Tang, M. Plihal and D. L. Mills, Journal of Magnetism and Magnetic
Materials 187, 23 (1998); R. B. Muniz and D. L. Mills, Phys. Rev. B 66, 174417 (2002); A. T.
Costa, R. B. Muniz and D. L. Mills, Phys. Rev. B 68, 224435 (2003).
266For an explicit comparison between the predictions of the He isenberg model and that of an
itinerant electron description of the spin excitations, se e Fig. 4 and Fig. 5 of the theoretical
paper cited in ref. 4.
7S. Blgel, M. Weinert and P. H. Dederichs, Phys. Rev. Lett. 60, 1077 (1988).
8S. Heinze, M. Bode, A. Kubetzka, O. Pietzsch, X. Nie, S. Blgel , and R. Wiesendanger, Science
311, 1805 (2000).
9M. Bode, M. Heide, K. von Bergmann, P. Ferriani, S. Heinze, G. Bihlmayer, A. Kubetzka, O.
Pietzsch, S. Blgel and R. Wiesendanger, Nature 447, 190 (2007).
10P. Lederer and A.Blandin, Phil. Mag. 14, 363 (1966).
11See also A. T. Costa, R. B. Muniz and D. L. Mills, Phys. Rev. B 73, 214403 (2006).
12L. Udvardi and L. Szunyogh, Phys. Rev. Lett. 102, 207204 (2009).
13E. Abate and M. Asdente, Phys. Rev. 140, A1303 (1965).
14A. T. Costa, Roberto Bechara Muniz and D. L. Mills, Phys. Rev. B73, 054426 (2006).
15Peter Fulde and Alan Luther, Phys. Rev. 175, 337 (1968).
16See the discussion in the second paper cited in footnote 5.
17See Chapter 6 and Chapter 7 and Appendix C of Group Theory and Q uantum Mechanics, M.
Tinkham, (McGraw Hill, New York, 1964).
18See D. M. Edwards, page 114 of Moment Formation in Solids, Vol . 117 of NATO Advanced
Study Institute, Series B: Physics (Plenum, New York, 1984) , edited by W. J. L. Buyers.
19See the second and third papers cited in footnote 5.
20A. T. Costa, R. B. Muniz, J. X. Cao, R. Q. Wu and D. L. Mills, Phys . Rev. B 78, 054439
(2008).
21J. Prokop, W. X. Tang, Y. Zhang, I Tudosa, T. R. F. Peixoto, Kh. Zakeri, and J. Kirschner,
Phys. Rev. Lett. 102, 177206 (2009).
22W. X. Tang, Y. Zhang, I. Tudosa, J. Prokop, M. Etzkorn and J. Ki rschner, Phys. Rev. Lett.
99, 087202 (2007).
23Seethe discussion in A. T. Costa, R. B. Munizand D. L. Mills, I EEE Transactions on Magnetics
44, 1974 (2008).
24S. Krause, G. Herzog, T. Stapelfeldt, L. Berbil-Bautista, M . Bode E. Y. Vedmedenko, and R.
Wiesendanger, Phys. Rev. Lett. 103, 127202 (2009).
25M. Pratzer, H. J. Elmers, M. Bode, O. Pietzsch, A. Kubetzka an d R. Wiesendanger, Phys. Rev.
27Lett.88, 057201 (2002).
26N. Papanikolaou, R. Zeller and P. H. Dederichs, J. Phys. Cond ens. Matter 14, 2799 (2002).
27S. H. Vosko, L. Wilk and M. Nusair, Can. J. Phys. 58, 1200 (1980).
28A. I. Liechtenstein, M. I. Katsnelson, V. P. Antropov, and V. A. Gubanov, J. Magnetism and
Magnetic Materials 67, 65 (1987).
29P. R. Peduto, S. Frota-Pessa and M. S. Methfessel, Phys. Rev. B44, 13283 (1991).
30S. Frota-Pessˆ oa, Phys. Rev. B 46, 14570 (1992).
31R. Haydock, Solid State Physics, edited by H. Ehrenreich, F. Seitz and D. Turnbull, (Academic
Press, New York, 1980), vol. 35.
32O. K. Andersen, Phys. Rev. B 12, 3060 (1975).
33O. K. Andersen, O. Jepsen and D. Gltzel, Highlights of Conden sed Matter Theory, edited by
F. Bassani, F. Fumi, and M. P. Tosi, (North Holland, Amsterda m, 1985).
34U. von Barth and L. A. Hedin, J. Phys. C. 5, 1629 (1972).
35O. Eriksson, B. Johansson, R. C. Albers, A. M. Boring, and M. S . S. Brooks, Phys. Rev. B 42,
2707 (1990).
36N. Beer and D. Pettifor, The Electronic Structure of Complex Systems, (Plenum Press, New
York, 1984).
37M. Plihal and D. L. Mills, Phys. Rev. B 58, 14407 (1998).
38M. Heide, G. Bihlmayer and S. Blgel, Phys. Rev. B 78, 140403 (2008).
39Kh. Zakeri, Y. Zhang, J. Prokop, T. H. Chuang, N. Sakr, W. X. Ta ng and J. Kirschner, Phys.
Rev. Lett. 104, 137203 (2010).
40B. M. Hall andD. L. Mills, Phys. Rev. B 34, 8318 (1986); BurlM. Hall, D. L. Mills, Mohammed
H. Mohammed and L. L. Kesmodel, Phys. Rev. B 38, 5856 (1988).
41M. P. Gokhale, A. Ormeci and D. L. Mills, Phys. Rev. B 46, 8978 (1992).
42B. M. Hall, S. Y. Tong and D. L. Mills, Phys. Rev. Lett. 50, 1277 (1983).
43R. Urban, G. Woltersdorf, and B. Heinrich, Phys. Rev. Lett. 87, 217204 (2001).
44A. T. Costa, R. B. Muniz and D. L. Mills, Phys. Rev. B 73, 054426 (2006).
45A. Bergman, A. Taroni, L. Bergqvist, J. Hellsvik, B. Hrvarss on, and O. Eriksson, Phys. Rev. B
81, 144416 (2010).
28 |
1408.1838v3.Spin_Orbital_Order_Modified_by_Orbital_Dilution_in_Transition_Metal_Oxides__From_Spin_Defects_to_Frustrated_Spins_Polarizing_Host_Orbitals.pdf | Spin-Orbital Order Modied by Orbital Dilution in Transition Metal Oxides:
From Spin Defects to Frustrated Spins Polarizing Host Orbitals
Wojciech Brzezicki,1, 2Andrzej M. Ole s,3, 1and Mario Cuoco2
1Marian Smoluchowski Institute of Physics, Jagiellonian University,
prof. S. Lojasiewicza 11, PL-30348 Krak ow, Poland
2CNR-SPIN, IT-84084 Fisciano (SA), Italy, and
Dipartimento di Fisica \E. R. Caianiello", Universit a di Salerno, IT-84084 Fisciano (SA), Italy
3Max-Planck-Institut f ur Festk orperforschung, Heisenbergstrasse 1, D-70569 Stuttgart, Germany
(Dated: 24 December 2014)
We investigate the changes in spin and orbital patterns induced by magnetic transition metal
ions without an orbital degree of freedom doped in a strongly correlated insulator with spin-orbital
order. In this context we study the 3 dion substitution in 4 dtransition metal oxides in the case of
3d3doping at either 3 d2or 4d4sites which realizes orbital dilution in a Mott insulator. Although we
concentrate on this doping case as it is known experimentally and more challenging than other oxides
due to nite spin-orbit coupling, the conclusions are more general. We derive the eective 3 d 4d(or
3d 3d) superexchange in a Mott insulator with dierent ionic valencies, underlining the emerging
structure of the spin-orbital coupling between the impurity and the host sites and demonstrate
that it is qualitatively dierent from that encountered in the host itself. This derivation shows
that the interaction between the host and the impurity depends in a crucial way on the type of
doubly occupied t2gorbital. One nds that in some cases, due to the quench of the orbital degree
of freedom at the 3 dimpurity, the spin and orbital order within the host is drastically modied by
doping. The impurity acts either as a spin defect accompanied by an orbital vacancy in the spin-
orbital structure when the host-impurity coupling is weak, or it favors doubly occupied active orbitals
(orbital polarons) along the 3 d 4dbond leading to antiferromagnetic or ferromagnetic spin coupling.
This competition between dierent magnetic couplings leads to quite dierent ground states. In
particular, for the case of a nite and periodic 3 datom substitution, it leads to striped patterns either
with alternating ferromagnetic/antiferromagnetic domains or with islands of saturated ferromagnetic
order. We nd that magnetic frustration and spin degeneracy can be lifted by the quantum orbital
ips of the host but they are robust in special regions of the incommensurate phase diagram.
Orbital quantum
uctuations modify quantitatively spin-orbital order imposed by superexchange. In
contrast, the spin-orbit coupling can lead to anisotropic spin and orbital patterns along the symmetry
directions and cause a radical modication of the order imposed by the spin-orbital superexchange.
Our ndings are expected to be of importance for future theoretical understanding of experimental
results for 4 dtransition metal oxides doped with 3 d3ions. We suggest how the local or global
changes of the spin-orbital order induced by such impurities could be detected experimentally.
PACS numbers: 75.25.Dk, 03.65.Ud, 64.70.Tg, 75.30.Et
I. INTRODUCTION
The studies of strongly correlated electrons in transi-
tion metal oxides (TMOs) focus traditionally on 3 dma-
terials [1], mainly because of high-temperature super-
conductivity discovered in cuprates and more recently
in iron-pnictides, and because of colossal magnetoresis-
tance manganites. The competition of dierent and com-
plex types of order is ubiquitous in strongly correlated
TMOs mainly due to coupled spin-charge-orbital where
frustrated exchange competes with the kinetic energy of
charge carriers. The best known example is spin-charge
competition in cuprates, where spin, charge and super-
conducting orders intertwine [2] and stripe order emerges
in the normal phase as a compromise between the mag-
netic and kinetic energy [3, 4]. Remarkable evolution
of the stripe order under increasing doping is observed
[5] and could be reproduced by the theory based on the
extended Hubbard model [6]. Hole doping in cuprates
corresponds to the removal of the spin degree of free-dom. Similarly, hole doping in a simplest system with the
orbital order in d1conguration removes locally orbital
degrees of freedom and generates stripe phases which in-
volve orbital polarons [7]. It was predicted recently that
orbital domain walls in bilayer manganites should be par-
tially charged as a result of competition between orbital-
induced strain and Coulomb repulsion [8], which opens a
new route towards charge-orbital physics in TMOs. We
will show below that the stripe-like order may also occur
in doped spin-orbital systems. These systems are very
challenging and their doping leads to very complex and
yet unexplored spin-orbital-charge phenomena [9].
A prerequisite to the phenomena with spin-orbital-
charge coupled degrees of freedom is the understanding
of undoped systems [10], where the low-energy physics
and spin-orbital order are dictated by eective spin-
orbital superexchange [11{13] and compete with spin-
orbital quantum
uctuations [14{16]. Although ordered
states occur in many cases, the most intriguing are quan-
tum phases such as spin [17] or orbital [18] liquids. Re-arXiv:1408.1838v3 [cond-mat.str-el] 30 Jan 20152
cent experiments on a copper oxide Ba 3CuSb 2O9[19, 20]
have triggered renewed eorts in a fundamental search for
a quantum spin-orbital liquid [21{24], where spin-orbital
order is absent and electron spins are randomly choosing
orbitals which they occupy. A signature of strong quan-
tum eects in a spin-orbital system is a disordered state
which persists down to very low temperatures. A good
example of such a disordered spin-orbital liquid state is
as well FeSc 2S4which does not order in spite of nite
Curie-Weiss temperature CW= 45 K [25], but shows
instead signatures of quantum criticality [26, 27].
Spin-orbital interactions may be even more challenging
| for instance previous attempts to nd a spin-orbital
liquid in the Kugel-Khomskii model [14] or in LiNiO 2[28]
turned out to be unsuccessful. In fact, in the former case
certain types of exotic spin order arise as a consequence
of frustrated and entangled spin-orbital interactions [29,
30], and a spin-orbital entangled resonating valence bond
state was recently shown to be a quantum superposition
of strped spin-singlet covering on a square lattice [31]. In
contrast, spin and orbital superexchange have dierent
energy scales and orbital interactions in LiNiO 2are much
stronger and dominated by frustration [32]. Hence the
reasons behind the absence of magnetic long range order
are more subtle [33]. In all these cases orbital
uctuations
play a prominent role and spin-orbital entanglement [34]
determines the ground state.
The role of charge carriers in spin-orbital systems is
under very active investigation at present. In doped
La1 x(Sr,Ca)xMnO 3manganites several dierent types
of magnetic order compete with one another and occur
at increasing hole doping [35{37]. Undoped LaMnO 3
is an antiferromagnetic (AF) Mott insulator, with large
S= 2 spins for 3 d4ionic congurations of Mn3+ions
stabilized by Hund's exchange, coupled via the spin-
orbital superexchange due to egandt2gelectron exci-
tations [38]. The orbital egdegree of freedom is re-
moved by hole doping when Mn3+ions are generated,
and this requires careful modeling in the theory that
takes into account both 3 d4and 3d3electronic cong-
urations of Mn3+and Mn4+ions [39{44]. In fact, the
orbital order changes radically with increasing doping
in La 1 x(Sr,Ca)xMnO 3systems at the magnetic phase
transitions between dierent types of magnetic order [37],
as weel as at La 0:7Ca0:3MnO 3/BiFeO 3heterostructures,
where it oers a new route to enhancing multiferroic func-
tionality [45]. The double exchange mechanism [46] trig-
gers ferromagnetic (FM) metallic phase at sucient dop-
ing; in this phase the spin and orbital degrees of freedom
decouple and spin excitations are explained by the or-
bital liquid [47, 48]. Due to distinct magnetic and kinetic
energy scales, even low doping may suce for a drastic
change in the magnetic order, as observed in electron-
doped manganites [49].
A rather unique example of a spin-orbital system with
strongly
uctuating orbitals, as predicted in the theory
[50{52] and seen experimentally [53{55], are the per-
ovskite vanadates with competing spin-orbital order [56].In theset2gsystemsxyorbitals are lled by one electron
and orbital order of active fyz;zxgorbitals is strongly
in
uenced by doping with Ca (Sr) ions which replace Y
(La) ones in YVO 3(LaVO 3). In this case nite spin-
orbit coupling modies the spin-orbital phase diagram
[57]. In addition, the AF order switches easily from the
G-type AF (G-AF) toC-type AF (C-AF) order in the
presence of charge defects in Y 1 xCaxVO3. Already at
lowx'0:02 doping the spin-orbital order changes and
spectral weight is generated within the Mott-Hubbard
gap [58]. Although one might imagine that the orbital
degree of freedom is thereby removed, a closer inspec-
tion shows that this is not the case as the orbitals are
polarized by charge defects [59] and readjust near them
[60]. Removing the orbital degree of freedom in vana-
dates would be only possible by electron doping generat-
ing instead d3ionic congurations, but such a doping by
charge defects would be very dierent from the doping
by transition metal ions of the same valence considered
below.
Also in 4dmaterials spin-orbital physics plays a role
[61], as for instance in Ca 2 xSrxRuO 4systems with Ru4+
ions in 4d4conguration [62{66]. Recently it has been
shown that unconventional magnetism is possible for
Ru4+and similar ions where spin-orbit coupling plaus a
role [67, 68]. Surprisingly, these systems are not similar
to manganites but to vanadates where one nds as well
ions with active t2gorbitals. In the case of ruthenates
thet4
2gRu4+ions have low S= 1 spin as the splitting
between the t2gandeglevels is large. Thus the undoped
Ca2RuO 4is a hole analogue of a vanadate [50, 51], with
t2gorbital degree of freedom and S= 1 spin per site in
both cases. This gives new opportunities to investigate
spin-orbital entangled states in t2gsystem, observed re-
cently by angle resolved photoemission [69].
Here we focus on a novel and very dierent doping
from all those considered above, namely on a substitu-
tional doping by other magnetic ions in a plane built
by transition metal and oxygen ions, for instance in the
(a;b) plane of a monolayer or in perovskite ruthenates
or vanadates. In this study we are interested primarily
in doping of a TMO with t2gorbital degrees of freedom,
where doped magnetic ions have no orbital degree of free-
dom and realize orbital dilution . In addition, we deal with
the simpler case of 3 ddoped ions where we can neglect
spin-orbit interaction which should not be ignored for 4 d
ions. We emphasize that in contrast to manganites where
holes within egorbitals participate in transport and are
responsible for the colossal magnetoresitance, such doped
hole are immobile due to the ionic potential at 3 dsites
and form defects in spin-orbital order of a Mott insulator.
We encounter here a dierent situation from the dilution
eects in the 2D egorbital system considered so far [70]
as we deel with magnetic ions at doped sites. It is chal-
lenging to investigate how such impurities modify locally
or globally spin-orbital order of the host.
The doping which realizes this paradigm is by either
Mn4+or Cr3+ions with large S= 3=2 spins stabilized by3
FIG. 1. (a) Schematic view of the orbital dilution when the
3d3ion with no orbital degree of freedom and spin S= 3=2
substitutes 4 d4one with spin S= 1 on a bond having specic
spin and orbital character in the host (gray arrows). Spins
are shown by red arrows and doubly occupied t2gorbitals
(doublons) are shown by green symbols for aandcorbitals,
respectively. (b) If an inactive orbital along the bond is re-
moved by doping, the total spin exchange is AF. (c) On the
contrary, active orbitals at the host site can lead to either FM
(top) or AF (bottom) exchange coupling, depending on the
energy levels mismatch and dierence in the Coulomb cou-
plings between the impurity and the host. We show the case
when the host site is unchanged in the doping process.
Hund's exchange, and orbital dilution occurs either in a
TMO with d2ionic conguration as in the vanadium per-
ovskites, or in 4 dMott insulators as in ruthenates. It has
been shown that dilute Cr doping for Ru reduces the tem-
perature of the orthorhombic distortion and induces FM
behavior in Ca 2Ru1 xCrxO4(with 0<x< 0:13) [71]. It
also induces surprising negative volume thermal expan-
sion via spin-orbital order. Such defects, on one hand,
can weaken the spin-orbital coupling in the host, but on
the other hand, may open a new channel of interaction
between the spin and orbital degree of freedom through
the host-impurity exchange, see Fig. 1. Consequences of
such doping are yet unexplored and are expected to open
a new route in the research on strongly correlated oxides.
The physical example for the present theory are the in-
sulating phases of 3 d 4dhybrid structures, where doping
happens at d4transition metal sites, and the value of the
spin is locally changed from S= 1 toS= 3=2. As a
demonstration of the highly nontrivial physics emerging
abFIG. 2. Schematic view of C-AF spin order coexisting with
G-AO orbital order in the ( a;b) plane of an undoped Mott
insulator with 4 d4ionic congurations. Spins are shown by
arrows while doubly occupied xyandyzorbitals (canda
doublons, see text) form a checkerboard pattern. Equivalent
spin-orbital order is realized for V3+ions in (b;c) planes of
LaVO 3[56], with orbitals standing for empty orbitals (holes).
in 3d 4doxides, remarkable eects have already been
observed, for instance, when Ru ions are replaced by Mn,
Ti, Cr or other 3 delements. The role of Mn doping in
the SrRuO Ruddlesden-Popper series is strongly linked
to the dimensionality through the number nof RuO 2
layers in the unit cell. The Mn doping of the SrRuO 3
cubic member drives the system from the itinerant FM
state to an insulating AF conguration in a continuous
way via a possible unconventional quantum phase tran-
sition [72]. Doping by Mn ions in Sr 3Ru2O7leads to
a metal-to-insulator transition and AF long-range order
for more than 5% Mn concentration [73]. Subtle orbital
rearrangement can occur at the Mn site, as for instance
the inversion of the crystal eld in the egsector observed
via x-ray absorption spectroscopy [74]. Neutron scatter-
ing studies indicate the occurrence of an unusual E-type
antiferromagnetism in doped systems (planar order with
FM zigzag chains with AF order between them) with mo-
ments aligned along the caxis within a single bilayer [75].
Furthermore, the more extended 4 dorbitals would a
priori suggest a weaker correlation than in 3 dTMOs due
to a reduced ratio between the intraatomic Coulomb in-
teraction and the electron bandwidth. Nevertheless, the
(eective)d-bandwidth is reduced by the changes in the
3d-2p-3dbond angles in distorted structures which typi-
cally arise in these materials. This brings these systems
on the verge of a metal-insulator transition [76], or even
into the Mott insulating state with spin-orbital order, see
Fig. 2. Hence, not only 4 dmaterials share common fea-
tures with 3 dsystems, but are also richer due to their sen-
sitivity to the lattice structure and to relativistic eects
due to larger spin-orbit [77] or other magneto-crystalline
couplings.
To simplify the analysis we assume that onsite
Coulomb interactions are so strong that charge degrees of
freedom are projected out, and only virtual charge trans-
fer can occur between 3 dand 4dions via the oxygen lig-4
ands. For convenience, we dene the orbital degree of
freedom as a doublon (double occupancy) in the t4
2gcon-
guration. The above 3 ddoping leads then eectively to
the removal of a doublon in one of t2gorbitals which we
label asfa;b;cg(this notation is introduced in Ref. [16]
and explained below) and to replacing it by a t3
2gion.
To our knowledge, this is the only example of remov-
ing the orbital degree of freedom in t2gmanifold realized
so far and below we investigate possible consequences of
this phenomenon. Another possibility of orbital dilution
which awaits experimental realization would occur when
at2gdegree of freedom is removed by replacing a d2ion
by ad3one, as for instance by Cr3+doping in a vanadate
| here a doublon is an empty t2gorbital, i.e., lled by
two holes.
Before presenting the details of the quantitative anal-
ysis, let us concentrate of the main idea of the superex-
change modied by doping in a spin-orbital system. The
d3ions have singly occupied all three t2gorbitals and
S= 3=2 spins due to Hund's exchange. While a pair
ofd3ions, e.g. in SrMnO 3, is coupled by AF superex-
change [48], the superexchange for the d3 d4bond has
a rather rich structure and may also be FM. The spin
exchange depends then on whether the orbital degree of
freedom is active and participates in charge excitations
along a considered bond or electrons of the doublon can-
not move along this bond due to the symmetry of t2g
orbital, as explained in Fig. 1. This qualitative dier-
ence to systems without active orbital degrees of freedom
is investigated in detail in Sec. II.
The main outcomes of our analysis are: (i) the de-
termination of the eective spin-orbital exchange Hamil-
tonian describing the low-energy sector for the 3 d 4d
hybrid structure, (ii) establishing that a 3 d3impurity
without an orbital degree of freedom modies the orbital
order in the 4 d4host, (iii) providing the detailed way how
the microscopic spin-orbital order within the 4 d4host is
modied around the 3 d3impurity, and (iv) suggesting
possible spin-orbital patterns that arise due to periodic
and nite substitution (doping) of 4 datoms in the host
by 3dones. The emerging physical scenario is that the
3dimpurity acts as an orbital vacancy when the host-
impurity coupling is weak and as an orbital polarizer of
the bonds active t2gdoublon congurations when it is
strongly coupled to the host. The tendency to polarize
the host orbitals around the impurity turns out to be ro-
bust and independent of spin conguration. Otherwise,
it is the resulting orbital arrangement around the impu-
rity and the strength of Hund's coupling at the impurity
that set the character of the host-impurity magnetic ex-
change.
The remaining of the paper is organized as follows. In
Sec. II we introduce the eective model describing the
spin-orbital superexchange at the 3 d 4dbonds which
serves to investigate the changes of spin and orbital or-
der around individual impurities and at nite doping.
We arrive at a rather general formulation which empha-
sizes the impurity orbital degree of freedom, being a dou-blon, and present some technical details of the deriva-
tion in Appendix A. The strategy we adopt is to analyze
rst the ground state properties of a single 3 d3impu-
rity surrounded by 4 d4atoms by investigating how the
spin-orbital pattern in the host may be modied at the
nearest neighbor (NN) sites to the 3 datom. This study
is performed for dierent spin-orbital patterns of the 4 d
host with special emphasis on the alternating FM chains
(C-AF order) which coexist with G-type alternating or-
bital (G-AO) order, see Fig. 2. We address the impurity
problem within the classical approximation in Sec. III A.
As explained in Sec. III B, there are two nonequivalent
cases which depend on the precise modication of the or-
bital order by the 3 dimpurity, doped either to replace a
doublon in aorbital (Sec. III C) or the one in corbital
(Sec. III D).
Starting from the single impurity solution we next ad-
dress periodic arrangements of 3 datoms at dierent con-
centrations. We demonstrate that the spin-orbital or-
der in the host can be radically changed by the presence
of impurities, leading to striped patterns with alternat-
ing FM/AF domains and islands of fully FM states. In
Sec. IV A we consider the modications of spin-orbital
order which arise at periodic doping with macroscopic
concentration. Here we limit ourselves to two represen-
tative cases: (i) commensurate x= 1=8 doping in Sec.
IV B, and (ii) two doping levels x= 1=5 andx= 1=9 be-
ing incommensurate with underlying two-sublattice order
(Fig. 2) which implies simultaneous doping at two sub-
lattices, i.e., at both aandcdoublon sites, as presented
in Secs. IV C and IV D. Finally, in Sec. V A we inves-
tigate the modications of the classical phase diagram
induced by quantum
uctuations, and in Secs. V B and
V C we discuss representative results obtained for nite
spin-orbit coupling (calculation details of the treatment
of spin-orbit interaction are presented in Appendix B).
The paper is concluded by a general discussion of possi-
ble emerging scenarios for the 3 d3impurities in 4 d4host,
a summary of the main results and perspective of future
experimental investigations of orbital dilution in Sec. VI.
II. THE SPIN-ORBITAL MODEL
In this Section we consider a 3 dimpurity in a strongly
correlated 4 dTMO and derive the eective 3 d3 4d4
spin-orbital superexchange. It follows from the coupling
between 3dand 4dorbitals via oxygen 2 porbitals due to
thep dhybridization. In a strongly correlated system
it suces to concentrate on a pair of atoms forming a
bondhiji, as the eective interactions are generated by
charge excitations d4
id4
jd5
id3
jalong a single bond [12].
In the reference 4 dhost both atoms on the bond hijiare
equivalent and one considers,
H(i;j) =Ht(i;j) +Hint(i) +Hint(j): (1)
The Coulomb interaction Hint(i) is local at site iand we
describe it by the degenerate Hubbard model [80], see5
below.
We implement a strict rule that the hopping within
thet2gsector is allowed in a TMO only between two
neighboring orbitals of the same symmetry which are ac-
tive along the bond direction [15, 78, 79], and neglect
the interorbital processes originating from the octahe-
dral distortions such as rotation or tilting. Indeed, in
ideal undistorted (perovskite or square lattice) geometry
the orbital
avor is conserved as long as the spin-orbit
coupling may be neglected. The interorbital hopping ele-
ments are smaller by at least one order of magnitude and
may be treated as corrections in cases where distortions
play a role to the overall scenario established below.
The kinetic energy for a representative 3 d-2p-4dbond,
i.e., after projecting out the oxygen degrees of freedom,
is given by the hopping in the host /thbetween sites i
andj,
Ht(i;j) = thX
(
);
dy
idj+dy
jdi
:(2)
Heredy
iare the electron creation operators at site iin
the spin-orbital state ( ). The bondhijipoints along
one of the two crystallographic directions,
=a;b, in
the two-dimensional (2D) square lattice. Without distor-
tions, only two out of three t2gorbitals are active along
each bondh12iand contribute to Ht(i;j), while the third
orbital lies in the plane perpendicular to the
axis and
thus the hopping via oxygen is forbidden by symmetry.
This motivates a convenient notation as follows [15],
jaijyzi;jbijxzi;jcijxyi; (3)
with thet2gorbital inactive along a given direction
2
fa;b;cglabeled by the index
. We consider a 2D square
lattice with transition metal ions connected via oxygen
orbitals as in a RuO 2(a;b) plane of Ca 2RuO 4(SrRuO 3).
In this casejai(jbi) orbitals are active along the b(a)
axis, whilejciorbitals are active along both a;baxes.
To derive the superexchange in a Mott insulator, it
is sucient to consider a bond which connects nearest
neighbor sites,hijih 12i. Below we consider a bond
between an impurity site i= 1 occupied by a 3 dion and
a neighboring host 4 dion at sitej= 2. The Hamiltonian
for this bond can be then expressed in the following form,
H(1;2) =Ht(1;2) +Hint(1) +Hint(2) +Hion(2):(4)
The total Hamiltonian contains the kinetic energy term
Ht(1;2) describing the electron charge transfer via oxy-
gen orbitals, the onsite interaction terms Hint(m) for the
3d(4d) ion at site m= 1;2, and the local potential of the
4datom,Hion(2), which takes into account the mismatch
of the energy level structure between the two (4 dand 3d)
atomic species and prevents valence
uctuations when
the host is doped, even in the absence of local Coulomb
interaction.
The kinetic energy in Eq. (4) is given by,
Ht(1;2) = tX
(
);
dy
1d2+dy
2d1
;(5)wheredy
mis the electron creation operator at site m=
1;2 in the spin-orbital state ( ). The bondh12ipoints
along one of the two crystallographic directions,
=a;b,
and again the orbital
avor is conserved [15, 78, 79].
The Coulomb interaction on an atom at site m= 1;2
depends on two parameters [80]: (i) intraorbital Coulomb
repulsionUm, and (ii) Hund's exchange JH
m. The label m
stands for the ion and distinguishes between these terms
at the 3dand 4dion, respectively. The interaction is
expressed in the form,
Hint(m) =UmX
nm"nm# 2JH
mX
<~Sm~Sm
+
Um 5
2JH
mX
<
0nmnm0
+JH
mX
6=dy
m"dy
m#dm#dm": (6)
The terms standing in the rst line of Eq. (6) con-
tribute to the magnetic instabilities in degenerate Hub-
bard model [80] and decide about spin order, both in an
itinerant system and in a Mott insulator. The remaining
terms contribute to the multiplet structure and are of im-
portance for the correct derivation of the superexchange
which follows from charge excitations, see below.
Finally, we include a local potential on the 4 datom
which encodes the energy mismatch between the host
and the impurity orbitals close to the Fermi level and
prevents valence
uctuations on the 4 dion due to the 3 d
doping. This term has the following general structure,
Hion(2) =Ie
2
4 X
;n2!2
; (7)
with=a;b;c .
The eective Hamiltonian for the low energy processes
is derived from H(1;2) (4) by a second order expansion
for charge excitations generated by Ht(1;2), and treating
the remaining part of H(1;2) as an unperturbed Hamil-
tonian. We are basically interested in virtual charge ex-
citations in the manifold of degenerate ground states of
a pair of 3dand 4datoms on a bond, see Fig. 3. These
quantum states are labeled as
ek
1
withk= 1;:::; 4 and
fep
2gwithp= 1;:::; 9 and their number follows from the
solution of the onsite quantum problem for the Hamilto-
nianHint(i). For the 3 datom the relevant states can be
classied according to the four components of the total
spinS1= 3=2 for the 3dimpurity atom at site m= 1,
three components of S2= 1 spin for the 4 dhost atom
at sitem= 2 and for the three dierent positions of the
double occupied orbital (doublon). Thus, the eective
Hamiltonian will contain spin products ( ~S1~S2) between
spin operators dened as,
~Sm=1
2X
dy
m
~ dm
; (8)6
3d atom 4d atomabc
eI2site 1 site 2
FIG. 3. Schematic representation of one conguration be-
longing to the manifold of 36 degenerate ground states for a
representative 3 d 4dbondh12ias given by the local Coulomb
Hamiltonian Hint(m) (6) withm= 1;2. The dominant ex-
change processes considered here are those that move one of
the four electrons on the 4 datom to the 3 dneighbor and back.
The stability of the 3 d3-4d4charge congurations is provided
by the local potential energy Ie
2, see Eq. (7).
form= 1;2 sites and the operator of the doublon posi-
tion at site m= 2,
D(
)
2=
dy
2
"d2
"
dy
2
#d2
#
: (9)
The doublon operator identies the orbital
within the
t2gmanifold of the 4 dion with a double occupancy (oc-
cupied by the doublon) and stands in what follows for
the orbital degree of freedom. It is worth noting that the
hopping (5) does not change the orbital
avor thus we
expect that the resulting Hamiltonian is diagonal in the
orbital degrees of freedom with only D(
)
2operators.
Following the standard second order perturbation ex-
pansion for spin-orbital systems [12], we can write the
matrix elements of the low energy exchange Hamiltonian,
H(
)
J(i;j), for a bondh12ik
along the
axis as follows,
ek
1;el
2H(
)
J(1;2)ek0
1;el0
2
= X
n1;n21
"n1+"n2
ek
1;el
2Ht(1;2)n1;n2
n1;n2Ht(1;2)ek0
1;el0
2
;(10)
with"nm=En;m E0;mbeing the excitation energies for
atoms at site m= 1;2 with respect to the unperturbed
ground state. The superexchange Hamiltonian H(
)
J(1;2)
for a bond along
can be expressed in a matrix form
by a 3636 matrix, with dependence on Um,JH
m, and
Ieelements. There are two types of charge excitations:
(i)d3
1d4
2d4
1d3
2one which creates a doublon at the 3 d
impurity, and (ii) d3
1d4
2d2
1d5
2one which adds another
doublon at the 4 dhost site in the intermediate state.
The second type of excitations involves more doubly oc-
cupied orbitals and has much larger excitation energy. It
is therefore only a small correction to the leading term
(i), as we discuss in Appendix A.
Similar as in the case of doped manganites [48], the
dominant contribution to the eective low-energy spin-
orbital Hamiltonian for the 3 d 4dbond stems from thed3
1d4
2d4
1d3
2charge excitations, as they do not involve
an extra double occupancy and the Coulomb energy U2.
The 3d3
14d4
23d4
14d3
2charge excitations can be analyzed
in a similar way as the 3 d3
i3d4
j3d4
i3d3
jones for anhiji
bond in doped manganites [48]. In both cases the total
number of doubly occupied orbitals does not change, so
the main contributions come due to Hund's exchange. In
the present case, one more parameter plays a role,
=Ie+ 3(U1 U2) 4(JH
1 JH
2); (11)
which stands for the mismatch potential energy (7) renor-
malized by the onsite Coulomb interactions fUmgand by
Hund's exchange fJH
mg. On a general ground we expect
to be a positive quantity, since the repulsion Umshould
be larger for smaller 3 dshells than for the 4 dones and
Umis the largest energy scale in the problem.
Let us have a closer view on this dominant contribu-
tion of the eective low-energy spin-orbital Hamiltonian
for the 3d 4dbond, given by Eq. (A2). For the anal-
ysis performed below and the clarity of our presentation
it is convenient to introduce some scaled parameters re-
lated to the interactions within the host and between the
host and the impurity. For this purpose we employ the
exchange couplings JimpandJhost,
Jimp=t2
4; (12)
Jhost=4t2
h
U2; (13)
which follow from the virtual charge excitations gener-
ated by the kinetic energy, see Eqs. (2) and (5). We use
their ratio to investigate the in
uence of the impurity on
the spin-orbital order in the host. Here this the hopping
amplitude between two t2gorbitals at NN 4 datoms,JH
2
andU2refer to the host, and (11) is the renormalized
ionization energy of the 3 d 4dbonds. The results de-
pend as well on Hund's exchange element for the impurity
and on the one at host atoms,
imp=JH
1
; (14)
host=JH
2
U2; (15)
Note that the ratio introduced for the impurity, imp(14),
has here a dierent meaning from Hund's exchange used
here for the host, host(15), which cannot be too large
by construction, i.e., host<1=3.
With the parametrization introduced above, the dom-
inant term in the impurity-host Hamiltonian for the im-
purity spin ~Siinteracting with the neighboring host spins
f~Sjgatj2N(i), deduced fromH(
)
3d 4d(1;2) Eq. (A2),
can be written in a rather compact form as follows
H3d 4d(i)'X
;j2N(i)n
JS(D(
)
j)(~Si~Sj) +EDD(
)
jo
;
(16)7
0.0 0.2 0.4 0.6 0.8 1.0
ηimp-0.20.00.20.40.6JS/Jimp, ED/JimpED
JS(Dγ= 1)
JS(Dγ= 0)22
FIG. 4. Evolution of the spin exchange JS(D(
)
2) and the
doublon energy ED, both given in Eq. (16) for increasing
Hund's exchange impat the impurity.
where the orbital (doublon) dependent spin couplings
JS(D(
)
j) and the doublon energy EDdepend on imp.
The evolution of the exchange couplings are shown in
Fig. 4. We note that the dominant energy scale is E
D,
so for a single 3 d 4dbond the doublon will avoid occupy-
ing the inactive (
) orbital and the spins will couple with
JS(D(
)
j= 0) which can be either AF if imp.0:43 or
FM ifimp>0:43. Thus the spins at imp=c
imp'0:43
will decouple according to the H(
)
J(i;j) exchange.
Let us conclude this Section by writing the complete
superexchange Hamiltonian,
H=H3d 4d+H4d 4d+Hso; (17)
whereH3d 4dP
iH3d 4d(i) includes all the 3 d 4d
bonds around impurities, H4d 4dstands for the the ef-
fective spin-orbital Hamiltonian for the 4 dhost bonds,
andHsois the spin-orbit interaction in the host. The
former term we explain below, while the latter one is
dened in Sec. V B, where we analyze the quantum cor-
rections and the consequences of spin-orbit interaction.
The superexchange in the host for the bonds hijialong
the
=a;baxes [81],
H4d 4d=JhostX
hijik
n
J(
)
ij(~Si~Sj+ 1) +K(
)
ijo
;(18)
depends on J(
)
ijandK(
)
ijoperators acting only in the
orbital space. They are expressed in terms of the pseu-
dospin operators dened in the orbital subspace spanned
by the two orbital
avors active along a given direction
, i.e.,
J(
)
ij=1
2(2r1+ 1)
~ i~ j(
) 1
2r2
z
iz
j(
)
+1
8
ninj(
)(2r1 r2+1) 1
4r1
ni+nj(
);(19)K(
)
ij=r1(~ i~ j)(
)+r2
z
iz
j(
)+1
4(r1+r2)
ninj(
)
1
4(r1+ 1)
ni+nj(
): (20)
with
r1=host
1 3host; r 2=host
1 + 2host; (21)
standing for the multiplet structure in charge excitations,
and the orbital operators f~ (
)
i;n(
)
igthat for the
=c
axis take the form:
~ (c)
i=1
2
ay
iby
i
~
aibi|; (22)
n(c)
i=ay
iai+by
ibi: (23)
For the directions
=a;bin the considered ( a;b) plane
one nds equivalent expressions by cyclic permutation of
the axis labelsfa;b;cgin the above formulas. This prob-
lem is isomorphic with the spin-orbital superexchange in
the vanadium perovskites [50, 51], where a hole in the
fa;bgdoublet plays an equivalent role to the doublon in
the present case. The operators fay
i;by
i;cy
igare the dou-
blon (hard core boson) creation operators in the orbital
=a;b;c , respectively, and they satisfy the local con-
straint,
ay
iai+by
ibi+cy
ici= 1; (24)
meaning that exactly onedoublon (9) occupies one of
the threet2gorbitals at each site i. These bosonic occu-
pation operators coincide with the previously used dou-
blon occupation operators D(
)
j, i.e.,D(
)
j=
y
j
jwith
=a;b;c . Below we follow rst the classical procedure
to determine the ground states of single impurities in Sec.
III, and at macroscopic doping in Sec IV.
III. SINGLE 3d IMPURITY IN 4d HOST
A. Classical treatment of the impurity problem
In this Section we describe the methodology that we
applied for the determination of the phase diagrams for a
single impurity reported below in Secs. III C, and next at
macroscopic doping, as presented in Sec. IV. Let us con-
sider rst the case of a single 3 dimpurity in the 4 dhost.
Since the interactions in the model Hamiltonian are only
eective ones between NN sites, it is sucient to study
the modication of the spin-orbital order around the im-
purity for a given spin-orbital conguration of the host
by investigating a cluster of L= 13 sites shown in Fig. 5.
We assume the C-AF spin order (FM chains coupled an-
tiferromagnetically) accompanied by G-AO order within
the host which is the spin-orbital order occurring for the
realistic parameters of a RuO 2plane [81], see Fig. 2. Such
a spin-orbital pattern turns out to be the most relevant
one when considering the competition between the host8
FIG. 5. Schematic top view of the cluster used to obtain
the phase diagrams of the 3 dimpurity within the 4 dhost in
an (a;b) plane. The impurity is at the central site i= 0
which belongs to the corbital host sublattice. For the outer
sites in this cluster the spin-orbital conguration is xed and
determined by the undoped 4 dhost (with spins and cor-
bitals shown here) having C-AF/G-AO order, see Fig. 2. For
the central i= 0 site the spin state and for the host sites
i= 1;:::; 4 the spin-orbital congurations are determined by
minimizing the energy of the cluster.
and the impurity as due to the AO order within the ( a;b)
plane. Other possible congurations with uniform orbital
order and AF spin pattern, e.g. G-AF order, will also be
considered in the discussion throughout the manuscript.
The sitesi= 1;2;3;4 inside the cluster in Fig. 5 have
active spin and orbital degrees of freedom while the im-
purity at site i= 0 has only spin degree of freedom. At
the remaining sites the spin-orbital conguration is as-
signed, following the order in the host, and it does not
change along the computation.
To determine the ground state we assume that the
spin-orbital degrees of freedom are treated as classical
variables. This implies that for the bonds between atoms
in the host we use the Hamiltonian (18) and neglect quan-
tum
uctuations, i.e., in the spin sector we keep only the
zth (Ising) spin components and in the orbital one only
the terms which are proportional to the doublon occupa-
tion numbers (9) and to the identity operators. Similarly,
for the impurity-host bonds we use the Hamiltonian Eq.
(16) by keeping only the zth projections of spin operators.
Since we do neglect the
uctuation of the spin amplitude
it is enough to consider only the maximal and minimal
values ofhSz
iifor spinS= 3=2 at the impurity sites and
S= 1 at the host atoms. With these assumptions we can
construct all the possible congurations by varying the
spin and orbital congurations at the sites from i= 1
toi= 4 in the cluster shown in Fig. 5. Note that the
outer ions in the cluster belong all to the same sublattice,
so two distinct cases have to be considered to probe all
the congurations. Since physically it is unlikely that asingle impurity will change the orbital order of the host
globally thus we will not compare the energies from these
two cases and analyze two classes of solutions separately,
see Sec. III B. Then, the lowest energy conguration in
each class provides the optimal spin-orbital pattern for
the NNs around the 3 dimpurity. In the case of degener-
ate classical states, the spin-orbital order is established
by including quantum
uctuations.
In the case of a periodic doping analyzed in Sec. IV,
we use a similar strategy in the computation. Taking
the most general formulation, we employ larger clusters
having both size and shape that depend on the impurity
distribution and on the spin-orbital order in the host.
For this purpose, the most natural choice is to search
for the minimum energy conguration in the elementary
unit cell that can reproduce the full lattice by a suitable
choice of the translation vectors. This is computation-
ally expensive but doable for a periodic distribution of
the impurities that is commensurate to the lattice be-
cause it yields a unit cell of relatively small size for dop-
ing around x= 0:1. Otherwise, for the incommensurate
doping the size of the unit cell can lead to a conguration
space of a dimension that impedes nding of the ground
state. This problem is computationally more demanding
and to avoid the comparison of all the energy congu-
rations, we have employed the Metropolis algorithm at
low temperature to achieve the optimal conguration it-
eratively along the convergence process. Note that this
approach is fully classical, meaning that the spins of the
host and impurity are treated as Ising variables and the
orbital
uctuations in the host's Hamiltonian Eq. (18)
are omitted. They will be addressed in Sec. V A.
B. Two nonequivalent 3ddoping cases
The single impurity problem is the key case to start
with because it shows how the short-range spin-orbital
correlations are modied around the 3 datom due to
the host and host-impurity interactions in Eq. (17).
The analysis is performed by xing the strength between
Hund's exchange and Coulomb interaction within the
host (6) at host= 0:1, and by allowing for a variation of
the ratio between the host-impurity interaction (16) and
the Coulomb coupling at the impurity site. The choice
ofhost= 0:1 is made here because this value is within
the physically relevant range for the case of the ruthe-
nium materials. Small variations of hostdo not aect
the obtained results qualitatively.
As we have already discussed in the model derivation,
the sign of the magnetic exchange between the impurity
and the host depends on the orbital occupation of the
4ddoublon around the 3 dimpurity. The main aspect
that controls the resulting magnetic conguration is then
given by the character of the doublon orbitals around
the impurity, depending on whether they are active or
inactive along the considered 3 d 4dbond. To explore
such a competition quantitatively we investigate G-AO9
order for the host with alternation of aandcdoublon
congurations accompanied by the C-AF spin pattern,
see Fig. 2. Note that the aorbitals are active only along
thebaxis, while the corbitals are active along the both
axes:aandb[79]. This state has the lowest energy for the
host in a wide range of parameters for Hund's exchange,
Coulomb element and crystal-eld potential [81].
Due to the specic orbital pattern of Fig. 2, the 3 dim-
purity can substitute one of two distinct 4 dsites which
are considered separately below, either with aor withc
orbital occupied by the doublon. Since the two 4 datoms
have nonequivalent surrounding orbitals, not always ac-
tive along the 3 d 4dbond, we expect that the result-
ing ground state will have a modied spin-orbital order.
Indeed, if the 3 datom replaces the 4 done with the dou-
blon in the aorbital, then all the 4 dneighboring sites
have active doublons along the connecting 3 d 4dbonds
because they are in the corbitals. On the contrary, the
substitution at the 4 dsite withcorbital doublon con-
guration leads to an impurity state with its neighbors
having both active and inactive doublons. Therefore, we
do expect a more intricate competition for the latter case
of an impurity occupying the 4 dsite withcorbital con-
guration. Indeed, this leads to frustrated host-impurity
interactions, as we show in Sec. III D.
C. Doping removing a doublon in aorbital
We start by considering the physical situation where
the 3dimpurity replaces a 4 dion with the doublon within
theaorbital. The ground state phase diagram and the
schematic view of the spin-orbital pattern are reported
in Fig. 6 in terms of the ratio Jimp=Jhost(14) and the
strength of Hund's exchange coupling imp(12) at the 3 d
site. There are three dierent ground states that appear
in the phase diagram. Taking into account the struc-
ture of the 3 d 4dspin-orbital exchange (16) we expect
that, in the regime where the host-impurity interaction
is greater than that in the host, the 4 dneighbors to the
impurity tend to favor the spin-orbital conguration set
by the 3d 4dexchange. In this case, since the orbitals
surrounding the 3 dsite already minimize the 3 d 4d
Hamiltonian, we expect that the optimal spin congura-
tion corresponds to the 4 dspins aligned either antifer-
romagnetically or ferromagnetically with respect to the
impurity 3dspin.
The neighbor spins are AF to the 3 dspin impurity
in the AFaphase, while the FM aphase is just obtained
from AFaby reversing the spin at the impurity, and hav-
ing all the 3 d 4dbonds FM. It is interesting to note that
due to the host-impurity interaction the C-AF spin pat-
tern of the host is modied in both the AF aand the
FMaground states. Another intermediate conguration
which emerges when the host-impurity exchange is weak
in the intermediate FS aphase where the impurity spin
is undetermined and its conguration in the initial C-
AF phase is degenerate with the one obtained after the
0.00.20.40.60.81.0ηimp
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Jimp/JhostAFaFMa
AFa FSa abFSa
?FIG. 6. Top | Phase diagram of the 3 dimpurity in the ( a;b)
plane with the C-AF/G-AO order in 4 dhost for the impurity
doped at the sublattice with an a-orbital doublon. Dierent
colors refer to local spin order around the impurity, AF and
FM, while FS indicates the intermediate regime of frustrated
impurity spin. Bottom | Schematic view of spin-orbital pat-
terns for the ground state congurations shown in the top
panel. The 3 datom is at the central site, the dotted frame
highlights the 4 dsites where the impurity induces a a spin
reversal. In the FS aphase the question mark stands for that
the frustrated impurity spin within the classical approach but
frustration is released by the quantum
uctuations of the NN
corbitals in the adirection resulting in small AF couplings
along theaaxis, and spins obey the C-AF order (small ar-
row). The labels FM aand AFarefer to the local spin order
around the 3 dimpurity site with respect to the host | these
states dier by spin inversion at the 3 datom site.
spin-inversion operation. This is a singular physical sit-
uation because the impurity does not select a specic
direction even if the surrounding host has a given spin-
orbital conguration. Such a degeneracy is clearly veri-
ed at the critical point c
imp'0:43 where the amplitude
of the 3d 4dcoupling vanishes when the doublon occu-
pies the active orbital. Interestingly, such a degenerate
conguration is also obtained at Jimp=Jhost<1 when the
host dominates and the spin conguration at the 4 dsites
around the impurity are basically determined by Jhost. In
this case, due to the C-AF spin order, always two bonds
are FM and other two have AF order, independently of
the spin orientation at the 3 dimpurity. This implies that10
orbital
fliporbital
flip0(a) (b)
FIG. 7. Schematic view of the two types of orbital bonds
found in the 4 dhost: (a) an active bond with respect to
orbital
ips, (
+
i
j+H:c:), and (b) an inactive bond, where
orbital
uctuations are blocked by the orbital symmetry |
here the orbitals are static and only Ising terms contribute to
the ground state energy.
both FM or AF couplings along the 3 d 4dbonds per-
fectly balance each other which results in the degenerate
FSaphase.
It is worth pointing out that there is a quite large re-
gion of the phase diagram where the FS astate is stabi-
lized and the spin-orbital order of the host is not aected
by doping with the possibility of having large degeneracy
in the spin conguration of the impurities. On the other
hand, by inspecting the corbitals around the impurity
(Fig. 6) from the point of view of the full host's Hamilto-
nian Eq. (18) with orbital
ips included, (
+
i
j+H:c:),
one can easily nd out that the frustration of the impu-
rity spin can be released by quantum orbital
uctuations.
Note that the corbitals around the impurity in the a
(b) direction have quite dierent surroundings. The ones
along theaaxis are connected by two active bonds along
thebaxis with orbitals a, as in Fig. 7(a), while the
ones alongbare connected with only one activeaorbital
along the same baxis. This means that in the perturba-
tive expansion the orbital
ips will contribute only along
thebbonds (for the present G-AO order) and admix the
aorbital character to corbitals along them, while such
processes will be blocked for the bonds along the aaxis,
as also forborbitals along the baxis, see Fig. 7(b).
This fundamental dierence can be easily included in
the host-impurity bond in the mean-eld manner by set-
tinghDib;
i= 0 for the bonds along the baxis and
0<hDia;
i<1 for the bonds along the aaxis. Then
one can easily check that for the impurity spin point-
ing downwards we get the energy contribution from the
spin-spin bonds which is given by E#=(host)hDia;
i,
and for the impurity spin pointing upwards we have
E"= (host)hDia;
i, with(host)>0. Thus, it is
clear that any admixture of the virtual orbital
ips in the
host's wave function polarize the impurity spin upwards
so that the C-AF order of the host will be restored.
D. Doping removing a doublon in corbital
Let us move to the case of the 3 datom replacing the
doublon at corbital. As anticipated above, this congu-
ration is more intricate because the orbitals surroundingthe impurity, as originated by the C-AF/AO order within
the host, lead to nonequivalent 3 d 4dbonds. There are
two bonds with the doublon occupying an inactive or-
bital (and has no hybridization with the t2gorbitals at
3d atom), and two remaining bonds with doublons in
activet2gorbitals.
Since the 3d 4dspin-orbital exchange depends on the
orbital polarization of 4 dsites we do expect a competition
which may modify signicantly the spin-orbital correla-
tions in the host. Indeed, one observes that three cong-
urations compete, denoted as AF1 c, AF2cand FMc, see
Fig. 8. In the regime where the host-impurity exchange
dominates the system tends to minimize the energy due
to the 3d 4dspin-orbital coupling and, thus, the orbitals
become polarized in the active congurations compatible
with theC-AF/G-AO pattern and the host-impurity spin
coupling is AF for imp0:43, while it is FM otherwise.
This region resembles orbital polarons in doped mangan-
ites [39, 42]. Also in this case, the orbital polarons arise
because they minimize the double exchange energy [46].
On the contrary, for weak spin-orbital coupling be-
tween the impurity and the host there is an interesting
cooperation between the 3 dand 4datoms. Since the
strength of the impurity-host coupling is not sucient to
polarize the orbitals at the 4 dsites, it is preferable to
have an orbital rearrangement to the conguration with
inactive orbitals on 3 d 4dbonds and spin
ips at 4 d
sites. In this way the spin-orbital exchange is optimized
in the host and also on the 3 d 4dbonds. The resulting
state has an AF coupling between the host and the im-
purity as it should when all the orbitals surrounding the
3datoms are inactive with respect to the bond direction.
This modication of the orbital conguration induces the
change in spin orientation. The double exchange bonds
(with inactive doublon orbitals) along the baxis are then
blocked and the total energy is lowered, in spite of the
frustrated spin-orbital exchange in the host. As a result,
the AF1cstate the spins surrounding the impurity are
aligned and antiparallel to the spin at the 3 dsite.
Concerning the host C-AF/G-AO order we note that
it is modied only along the direction where the FM cor-
relations develop and spin defects occur within the chain
doped by the 3 datom. The FM order is locally disturbed
by the 3ddefect antiferromagnetically coupled spins sur-
rounding it. Note that this phase is driven by the or-
bital vacancy as the host develops more favorable orbital
bonds to gain the energy in the absence of the orbital
degrees of freedom at the impurity. At the same time
the impurity-host bonds do not generate too big energy
losses as: (i) either impis so small that the loss due to
EDis compensated by the gain from the superexchange
/JS(D(
)
j= 1) (all these bonds are AF), see Fig. 4, or
(ii)Jimp=Jhostis small meaning that the overall energy
scale of the impurity-host exchange remains small. Inter-
estingly, if we compare the AF1 cwith the AF2 cground
states we observe that the disruption of the C-AF/G-
AO order is anisotropic and occurs either along the FM
chains in the AF1 cphase or perpendicular to the FM11
0.00.20.40.60.81.0ηimp
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Jimp/JhostAF1cFMc
AF2c
AF1c AF2c ab
FIG. 8. Top | Phase diagram of the 3 dimpurity in the
4dhost withC-AF/G-AO order and the impurity doped at
thecdoublon sublattice. Dierent colors refer to local spin
order around the impurity: AF c1, AFc2, and FM. Bottom |
Schematic view of spin-orbital patterns for the two AF ground
state congurations shown in the top panel; the FM cphase
diers from the AF c2 one only by spin inversion at the 3 d
atom. The 3 datom is at the central site and has no doublon
orbital, the frames highlight the spin-orbital defects caused
by the impurity. As in Fig. 6, the labels AF and FM refer to
the impurity spin orientation with respect to the neighboring
4dsites.
chains in the AF2 cphase. No spin frustration is found
here, in contrast to the FS aphase in the case of adoublon
doping, see Fig. 6.
Finally, we point out that a very similar phase diagram
can be obtained assuming that the host has the FM/ G-
AO order with aandborbitals alternating from site to
site. Such conguration can be stabilized by a distortion
that favors the out-of-plane orbitals. In this case there
is no dierence in doping at one or the other sublattice.
The main dierence is found in energy scales | for the
G-AF/C-AO order the diagram is similar to the one of
Fig. 8 if we rescale Jimpby half, which means that the
G-AF order is softer than the C-AF one. Note also that
in the peculiar AF1 cphase the impurity does not induce
any changes in the host for the FM/AO ordered host.
Thus we can safely conclude that the observed change in
the orbital order for the C-AF host in the AF1 cphase
is due to the presence of the corbitals which are notdirectional in the ( a;b) plane.
Summarizing, we have shown the complexity of local
spin-orbital order around t3
2gimpurities in a 4 t4
2ghost. It
is remarkable that such impurity spins not only modify
the spin-orbital order around them in a broad regime
of parameters, but also are frequently frustrated. This
highlights the importance of quantum eects beyond the
present classical approach which release frustration as we
show in Sec. V A.
IV. PERIODIC 3d DOPING IN 4d HOST
A. General remarks on nite doping
In this Section we analyze the spin-orbital patterns due
to a nite concentration xof 3dimpurities within the 4 d
host withC-AF/G-AO order, assuming that the 3 dim-
purities are distributed in a periodic way. The study is
performed for three representative doping distributions
| the rst one x= 1=8 is commensurate with the un-
derlying spin-orbital order and the other two are incom-
mensurate with respect to it, meaning that in such cases
doping at both aandcdoublon sites is imposed simulta-
neously.
As the impurities lead to local energy gains due to
3d 4dbonds surrounding them, we expect that the most
favorable situation is when they are isolated and have
maximal distances between one another. Therefore, we
selected the largest possible distances for the three dop-
ing levels used in our study: x= 1=8,x= 1=5, and
x= 1=9. This choice allows us to cover dierent regimes
of competition between the spin-orbital coupling within
the host and the 3 d 4dcoupling. While single impu-
rities may only change spin-orbital order locally, we use
here a high enough doping to investigate possible global
changes in spin-orbital order, i.e., whether they can oc-
cur in the respective parameter regime. The analysis is
performed as for a single impurity, by assuming the clas-
sical spin and orbital variables and by determining the
conguration with the lowest energy. For this analysis
we set the spatial distribution of the 3 datoms and we
determine the spin and orbital prole that minimizes the
energy.
B.C-AF phase with x= 1=8doping
We begin with the phase diagram obtained at x= 1=8
3ddoping, see Fig. 9. In the regime of strong impurity-
host coupling the 3 d 4dspin-orbital exchange deter-
mines the orbital and spin conguration of the 4 datoms
around the impurity. The most favorable state is when
the doublon occupies corbitals at the NN sites to the im-
purity. The spin correlations between the impurity and
the host are AF (FM), if the amplitude of impis below
(above)c
imp, leading to the AF aand the FM astates,
see Fig. 9. The AF astate has a striped-like prole with12
AF chains alternated by FM domains (consisting of three
chains) along the diagonal of the square lattice. Even if
the coupling between the impurity and the host is AF
for all the bonds in the AF astate, the overall congura-
tion has a residual magnetic moment originating by the
uncompensated spins and by the cooperation between
the spin-orbital exchange in the 4 dhost and that for the
3d 4dbonds. Interestingly, at the point where the domi-
nant 3d 4dexchange tends to zero (i.e., for imp'c
imp),
one nds a region of the FS aphase which is analogous
to the FSaphase found in Sec. III C for a single impu-
rity, see Fig. 6. Again the impurity spin is frustrated
in purely classical approach but this frustration is easily
released by the orbital
uctuations in the host so that
theC-AF order of the host can be restored. This state
is stable for the amplitude of impbeing close to c
imp.
The regime of small Jimp=Jhostratio is qualitatively
dierent | an orbital rearrangement around the impu-
rity takes place, with a preference to move the doublons
into the inactive orbitals along the 3 d 4dbonds. Such
orbital congurations favor the AF spin coupling at all
the 3d 4dbonds which is stabilized by the 4 d 4dsu-
perexchange [38]. This conguration is peculiar because
it generally breaks inversion and does not have any plane
of mirror symmetry. It is worth pointing out that the
original order in the 4 dhost is completely modied by
the small concentration of 3 dions and one nds that the
AF coupling between the 3 dimpurity and the 4 dhost
generally leads to patterns such as the AF cphase where
FM chains alternate with AF ones in the ( a;b) plane.
Another relevant issue is that the cooperation between
the host and impurity can lead to a fully polarized FM a
state. This implies that doping can release the orbital
frustration which was present in the host with the C-
AF/G-AO order.
C. Phase diagram for periodic x= 1=5doping
Next we consider doping x= 1=5 with a given periodic
spatial prole which concerns both doublon sublattices.
We investigate the 3 dspin impurities separated by the
translation vectors ~ u= (i;j) and~ v= (2; 1) (one can
show that for general periodic doping x,j~ uj2=x 1) so
there is a mismatch between the impurity periodicity and
the two-sublattice G-AO order in the host. One nds
that the present case, see Fig. 10, has similar general
structure of the phase diagram to the case of x= 1=8
(Fig. 9), with AF correlations dominating for implower
thanc
impand FM ones otherwise. Due to the specic
doping distribution there are more phases appearing in
the ground state phase diagram. For imp< c
impthe
most stable spin conguration is with the impurity cou-
pled antiferromagnetically to the host. This happens
both in the AF vacancy (AF v) and the AF polaronic
(AFp) ground states. The dierence between the two
AF states arises due to the orbital arrangement around
the impurity. For weak ratio of the impurity to the host
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Jimp/Jhost0.00.20.40.60.81.0ηimp
FMa
FSa
AFc AFaAFc
FSa AFa
ab
???
?FIG. 9. Top panel | Ground state diagram obtained for pe-
riodic 3ddopingx= 1=8. Dierent colors refer to local spin
order around the impurity: AF a, AFc, FSa, and FMa. Bot-
tom panel | Schematic view of the ground state congura-
tions within the four 8-site unit cells (indicated by blue dashed
lines) for the phases shown in the phase diagram. The ques-
tion marks in FS aphase indicate frustrated impurity spins
within the classical approach | the spin direction (small ar-
rows) is xed only by quantum
uctuations. The 3 datoms
are placed at the sites where orbitals are absent.
spin-orbital exchange, Jimp=Jhost, the orbitals around the
impurity are all inactive ones. On the contrary, in the
strong impurity-host coupling regime all the orbitals are
polarized to be in active (polaronic) states around the im-
purity. Both states have been found as AF1 cand AF2c
phase in the single impurity problem (Fig. 8).
More generally, for all phases the boundary given by
an approximate hyperbolic relation imp/J 1
impsepa-
rates the phases where the orbitals around impurities in13
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Jimp/Jhost0.00.20.40.60.81.0ηimp
AFp FMp
FMv
AFv FSvFS1p
FS2p
??
??
ab??
FIG. 10. Ground state diagram for x= 1=5 periodic concentration of 3 dimpurities (sites where orbitals are absent) with
schematic views of the ground state congurations obtained for the unit cell of 20 sites. Spin and orbital order are shown by
arrows and orbitals occupied by doublons; magnetic phases (AF, FS, and FM) are highlighted by dierent color. The question
marks in FS states (red circles) indicate frustrated impurity spins within the classical approach.
? ?(a) FSv (b) FS1p, FS2p
FIG. 11. Isotropic surrounding of the degenerate impurity
spins in the FS vand FSpphases in the case of x= 1=5 pe-
riodic doping (Fig. 10). Frames mark the clusters which are
not connected with orbitally active bonds.thec-orbital sublattice are all inactive (small imp) from
those where all the orbitals are active (large imp). The
inactive orbital around the impurity stabilize always the
AF coupling between the impurity spin and host spins
whereas the active orbitals can give either AF or FM
exchange depending on imp(hencec
imp, see Fig. 4).
Since the doping does not match the size of the elemen-
tary unit cell, the resulting ground states do not exhibit
specic symmetries in the spin-orbital pattern. They are
generally FM due to the uncompensated magnetic mo-
ments and the impurity feels screening by the presence
of the surrounding it host spins being antiparallel to the
impurity spin.
By increasing Hund's exchange coupling at the 3 dion
the system develops fully FM state in a large region of
the ground state diagram due to the possibility of suit-
able orbital polarization around the impurity. On the
other hand, in the limit where the impurity-host bonds14
ab0.0 0.5 1.0 1.5 2.0 2.5 3.0
Jimp/Jhost0.00.20.40.60.81.0ηimp
AFp FMp
FMv
AFv FSvFS1p
FS2p ??
??
??
FIG. 12. Ground state diagram for x= 1=9 periodic concentration of 3 dimpurities with schematic views of the ground state
congurations obtained for the unit cell of 36 sites. Spin order (AF, FS, and FM) is highlighted by dierent color. The question
marks in FS states (red squares) indicate frustrated impurity spins within the classical approach | the spin is xed here by
quantum
uctuations (small arrows). Doped 3 datoms are at the sites where orbitals are absent.
are weak, so either for imp'c
impand large enough
Jimp=Jhostso that all orbitals around the impurity are
active, or just for small Jimp=Jhostwe get the FS phases
where the impurity spin at the a-orbital sublattice is un-
determined in the present classical approach. This is a
similar situation to the one found in the FS aphase of a
single impurity problem and at x= 1=8 periodic doping,
see Figs. 6 and 9, but there it was still possible to iden-
tify the favored impurity spin polarization by considering
the orbital
ips in the host around the impurity.
However, the situation here is dierent as the host's
order is completely altered by doping and has became
isotropic, in contrast to the initial C-AF order (Fig. 2)
which breaks the planar symmetry between the aandb
direction. It was precisely this symmetry breaking that
favored one impurity spin polarization over the other one.
Here this mechanism is absent | one can easily checkthat the neighborhood of the corbitals surrounding im-
purities is completely equivalent in both directions (see
Fig. 11 for the view of these surroundings) so that the
orbital
ip argument is no longer applicable. This is a
peculiar situation in the classical approach and we indi-
cate frustration in spin direction by question marks in
Fig. 10.
In Fig. 11 we can see that both in FS vacancy (FS v)
and FS polaronic (FS p) phase the orbitals are grouped in
33 clusters and 22 plaquettes, respectively, that en-
circle the degenerate impurity spins. For the FS vphase
we can distinguish between two kind of plaquettes with
non-zero spin polarization diering by a global spin inver-
sion. In the case of FS pphases we observe four plaquettes
with zero spin polarization arranged in two pairs related
by a point re
ection with respect to the impurity site.
It is worthwhile to realize that these plaquettes are com-15
pletely disconnected in the orbital sector, i.e., there are
no orbitally active bonds connecting them (see Fig. 7 for
the pictorial denition of orbitally active bonds). This
means that quantum eects of purely orbital nature can
appear only at the short range, i.e., inside the plaquettes.
However, one can expect that if for some reason the two
degenerate spins in a single elementary cell will polarize
then they will also polarize in the same way in all the
other cells to favor long-range quantum
uctuations in
the spin sector related to the translational invariance of
the system.
D. Phase diagram for periodic x= 1=9doping
Finally we investigate low doping x= 1=9 with a given
periodic spatial prole, see Fig. 12. Here the impurities
are separated by the translation vectors ~ u= (0;3) and
~ v= (3;0). Once again there is a mismatch between the
periodic distribution of impurities and the host's two-
sublattice AO order, so we again call this doping incom-
mensurate as it also imposes doping at both doublon
sublattices. The ground state diagram presents gradu-
ally increasing tendency towards FM 3 d 4dbonds with
increasingimp, see Fig. 12. These polaronic bonds po-
larize as well the 4 d 4dbonds and one nds an almost
FM order in the FM pstate. Altogether, we have found
the same phases as at the higher doping of x= 1=5,
see Fig. 10, i.e., AF vand AFpat low values imp, FMv
and FMpin the regime of high imp, separated by the
regime of frustrated impurity spins which occur within
the phases: FS v, FS1p, and FS2p.
The dierence between the two AF (FM) states in Fig.
12 is due to the orbital arrangement around the impurity.
As for the other doping levels considered so far, x= 1=8
andx= 1=5, we nd neutral (inactive) orbitals around
3dimpurities in the regime of low Jimp=Jhostin AFvand
FMvphases which lead to spin defects within the 1D FM
chains in the C-AF spin order. A similar behavior was
reported for single impurities in the low doping regime in
Sec. III. This changes radically above the orbital tran-
sition for both types of local magnetic order, where the
orbitals reorient into the active ones. One nds that spin
orientations are then the same as those of their neigh-
boring 4datoms, with some similarities to those found
atx= 1=5, see Fig. 10.
Frustrated impurity spins occur in the crossover regime
between the AF and FM local order around impurities.
This follows from the local congurations around them,
which include two "-spins and two#-spins accompanied
bycorbitals at the NN 4 dsites. This frustration is easily
removed by quantum
uctuations and we suggest that
this happens again in the same way as for x= 1=8 doping,
as indicated by small arrows in the respective FS phases
shown in Fig. 12.V. QUANTUM EFFECTS BEYOND THE
CLASSICAL APPROACH
A. Spin-orbital quantum
uctuations
So far, we analyzed the ground states of 3 dimpurities
in the (a;b) plane of a 4 dsystem using the classical ap-
proach. Here we show that this classical picture may be
used as a guideline and is only quantitatively changed by
quantum
uctuations if the spin-orbit coupling is weak.
We start the analysis by considering the quantum prob-
lem in the absence of spin-orbit coupling (at = 0). The
orbital doublon densities,
N
X
i2hosthni
i; (25)
with
=a;b;c , and totalSzare conserved quantities and
thus good quantum numbers for a numerical simulation.
To determine the ground state congurations in the pa-
rameters space and the relevant correlation functions we
diagonalize exactly the Hamiltonian matrix (17) for the
cluster ofL= 8 sites by means of the Lanczos algorithm.
In Fig. 13(a) we report the resulting quantum phase
diagram for an 8-site cluster having one impurity and
assuming periodic boundary conditions, see Fig. 13(b).
This appears to be an optimal cluster conguration be-
cause it contains a number of sites and connectivities that
allows us to analyze separately the interplay between the
host-host and the host-impurity interactions and to sim-
ulate a physical situation when the interactions within
the host dominate over those between the host and the
impurity. Such a problem is a quantum analogue of the
single unit cell presented in Fig. 9 for x= 1=8 periodic
doping.
As a general feature that resembles the classical phase
diagram, we observe that there is a prevalent tendency
to have AF-like (FM-like) spin correlations between the
impurity and the host sites in the region of impbelow
(above) the critical point at c
imp'0:43 which separates
these two regimes, with intermediate congurations hav-
ing frustrated magnetic exchange. As we shall discuss
below it is the orbital degree of freedom that turns out
to be more aected by the quantum eects. Following
the notation used for the classical case, we distinguish
various quantum AF (QAF) ground states, i.e., QAF cn
(n= 1;2) and QAF an(n= 1;2), as well as a uni-
form quantum FM (QFM) conguration, i.e., QFM a, and
quantum frustrated one labeled as QFS a.
In order to visualize the main spin-orbital patterns con-
tributing to the quantum ground state it is convenient to
adopt a representation with arrows for the spin and el-
lipsoids for the orbital sector at any given host site. The
arrows stand for the on-site spin projection hSz
ii, with the
length being proportional to the amplitude. The length
scale for the arrows is the same for all the congurations.
Moreover, in order to describe the orbital character of
the ground state we employed a graphical representation
that makes use of an ellipsoid whose semi-axes fa;b;cg16
0.0 0.3 0.6 0.9 1.2 1.5
Jimp/Jhost0.000.250.500.751.00 ηimpQFMa
QFSa1
QAFc2 QAFc1QFSa2
QAFa1
QAFa2
12 8 34 5 673 2
6 41
71
71
7(a)
(b)
ab
FIG. 13. (a) Phase diagram for the quantum problem at zero spin-orbit simulated on the 8-site cluster in the presence of
one-impurity. Arrows and ellipsoids indicate the spin-orbital state at a given site i, while the shapes of ellipsoids re
ect the
orbital avarages: hay
iaii,hby
ibiiandhcy
icii(i.e., a circle in the plane perpendicular to the axis
implies 100% occupation of the
orbital
). (b) The periodic cluster of L= 8 sites used, with the orbital dilution (3 d3impurity) at site i= 8. The dotted lines
identify the basic unit cell adopted for the simulation with the same symmetries of the square lattice.
length are given by the average amplitude of the squared
angular momentum components f(Lx
i)2;(Ly
i)2;(Lz
i)2g, or
equivalently by the doublon occupation Eq. (9). For in-
stance, for a completely
at circle (degenerate ellipsoid)
lying in the plane perpendicular to the
axis only the
corresponding
orbital is occupied. On the other hand,
if the ellipsoid develops in all three directions fa;b;cgit
implies that more than one orbital is occupied and the
distribution can be anisotropic in general. If all the or-
bitals contribute equally, one nds an isotropic spherical
ellipsoid.
Due to the symmetry of the Hamiltonian, the phases
shown in the phase diagram of Fig. 13(a) can be
characterized by the quantum numbers for the z-th
spin projection, Sz, and the doublon orbital occupa-
tionN(25), (Sz;Na;Nb;Nc): QAFc1 ( 3:5;2;2;3),
QFSa2 ( 1:5;3;1;3), QAFa1 ( 5:5;1;3;3), QAFa2 and
QAFa2 ( 5:5;2;2;3), QFSa1 ( 0:5;3;0;4), and QFM a
( 8:5;2;1;4). Despite the irregular shape of the clus-
ter [Fig. 13(b)] there is also symmetry between the
aandbdirections. For this reason, the phases with
Na6=Nbcan be equivalently described either by the
set (Sz;Na;Nb;Nc) or (Sz;Nb;Na;Nc).
The outcome of the quantum analysis indicates thatthe spin patterns are quite robust as the spin congu-
rations of the phases QAF a, QAFc, QFSaand QFMa
are the analogues of the classical ones. The eects of
quantum
uctuations are more evident in the orbital sec-
tor where mixed orbital patterns occur if compared to
the classical case. In particular, orbital inactive states
around the impurity are softened by quantum
uctua-
tions and on some bonds we nd an orbital conguration
with a superposition of active and inactive states. The
unique AF states where the classical inactive scenario is
recovered corresponds to the QAF c1 and QAF c2 ones in
the regime of small imp. A small hybridization of active
and inactive orbitals along both the AF and FM bonds is
also observed around the impurity for the QFS aphases
as one can note by the shape of the ellipsoid at host sites.
Moreover, in the range of large impwhere the FM state
is stabilized, the orbital pattern around the impurity is
again like in the classical case.
A signicant orbital rearrangement is also obtained
within the host. We generally obtain an orbital pattern
that is slightly modied from the pure AO conguration
assumed in the classical case. The eect is dramatically
dierent in the regime of strong impurity-host coupling
(i.e., for large Jimp) with AF exchange (QAF a2) with the17
formation of an orbital liquid around the impurity and
within the host, with doublon occupation represented
by an almost isotropic shaped ellipsoid. Interestingly,
though with a dierent orbital arrangement, the QFS a1
and the QFS a2 states are the only ones where the C-AF
order of the host is recovered. For all the other phases
shown in the diagram of Fig. 13 the coupling between
the host and the impurity is generally leading to a uni-
form spin polarization with FM or AF coupling between
the host and the impurity depending on the strength of
the host-impurity coupling. Altogether, we conclude that
the classical spin patterns are only quantitatively modi-
ed and are robust with respect to quantum
uctuations.
B. Finite spin-orbit coupling
In this Section we analyze the quantum eects in the
spin and orbital order around the impurity in the pres-
ence of the spin-orbit coupling at the host d4sites. For
thet4
2gconguration the strong spin-orbit regime has
been considered recently by performing an expansion
around the atomic limit where the angular ~Liand spin
~Simomenta form a spin-orbit singlet for the amplitude
of the total angular momentum, ~Ji=~Li+~Si(i.e.,J= 0)
[67]. The instability towards an AF state starting from
theJ= 0 liquid has been obtained within the spin-wave
theory [68] for the low energy excitations emerging from
the spin-orbital exchange.
In the analysis presented here we proceed from the
limit of zero spin-orbit to investigate how the spin and
orbital order are gradually suppressed when approaching
theJ= 0 spin-orbit singlet state. This issue is addressed
by solving the full quantum Hamiltonian (17) exactly on
a cluster of L= 8 sites including the spin-orbital ex-
change for the host and that one derived for the host-
impurity coupling (17) as well as the spin-orbit term,
Hso=X
i2host~Li~Si: (26)
where the sum includes the ions of the 4 dhost and we
use the spin S= 1 and the angular momentum L= 1, as
in the ionic 4 d4congurations. Here is the spin-orbit
coupling constant at 4 dhost ions, and the components
of the orbital momentum ~LifLx
i;Ly
i;Lz
igare dened
as follows:
Lx
i=iX
(dy
i;xydi;xz dy
i;xzdi;xy);
Ly
i=iX
(dy
i;xydi;yz dy
i;yzdi;xy);
Lz
i=iX
(dy
i;xzdi;yz dy
i;yzdi;xz): (27)
To determine the ground state and the relevant corre-
lation functions we use again the Lanczos algorithm for
the cluster of L= 8 sites. Such an approach allows us tostudy the competition between the spin-orbital exchange
and the spin-orbit coupling on equal footing without any
simplifying approximation. Moreover, the cluster calcu-
lation permits to include the impurity in the host and
deal with the numerous degrees of freedom without mak-
ing approximations that would constrain the interplay of
the impurity-host versus host-host interactions.
Finite spin-orbit coupling signicantly modies the
symmetry properties of the problem. Instead of the
SU(2) spin invariance one has to deal with the rotational
invariance related to the total angular momentum per
site~Ji=~Li+~Si. Though the ~Li~Siterm in Eq. (26)
commutes with both total ~J2andJz, the full Hamilto-
nian for the host with impurities Eq. (17) has a reduced
symmetry because the spin sector is now linearly coupled
to the orbital which has only the cubic symmetry. Thus
the remaining symmetry is a cyclic permutation of the
fx;y;zgaxes.
Moreover,Jzis not a conserved quantity due to the or-
bital anisotropy of the spin-orbital exchange in the host
and the orbital character of the impurity-host coupling.
There one has a Z2symmetry associated with the parity
operator (-1)Jz. Hence, the ground state can be clas-
sied as even or odd with respect to the value of Jz.
This symmetry aspect can introduce a constraint on the
character of the ground state and on the impurity-host
coupling since the Jzvalue for the impurity is only due
to the spin projection while in the host it is due to the
combination of the orbital and spin projection. A direct
consequence is that the parity constraint together with
the unbalance between the spin at the host and the im-
purity sites leads to a nonvanishing total projection of
the spin and angular momentum with respect to a sym-
metry axis, e.g. the zthaxis. It is worth to note that a
xed parity for the impurity spin means that it prefers
to point in one direction rather than the other one which
is not the case for the host's spin and angular momen-
tum. Thus the presence on the impurity for a xed P
will give a nonzero polarization along the quantization
axis for every site of the system. Such a property holds
for any single impurity with a half-integer spin.
Another important consequence of the spin-orbit cou-
pling is that it introduces local quantum
uctuations in
the orbital sector even at the sites close to the impurity
where the orbital pattern is disturbed. The spin-orbit
term makes the on-site problem around the impurity ef-
fectively analogous to the Ising model in a transverse eld
for the orbital sector, with nontrivial spin-orbital entan-
glement [34] extending over the impurity neighborhood.
In Figs. 14 and 15 we report the schematic evolu-
tion of the ground state congurations for the cluster
ofL= 8 sites, with one-impurity and periodic boundary
conditions as a function of increasing spin-orbit coupling.
These patterns have been determined by taking into ac-
count the sign and the amplitude of the relevant spatial
dependent spin and orbital correlation functions. The
arrows associated to the spin degree of freedom can lie in
xyplane or out-of-plane (along z, chosen to be parallel18
QAFc1
QAFa1 QAFc2 QFMa
a
ba
ba
ba
b λ1
λ2
λ3
λ4
λ5
λ10λ1
λ2
λ3
λ4
λ7
λ10λ1
λ2
λ4
λ5
λ7
λ10λ1
λ2
λ3
λ4
λ7
λ10
FIG. 14. Evolution of the ground state congurations for the AF phases for selected increasing values of spin-orbit coupling
m, see Eq. (28). Arrows and ellipsoids indicate the spin-orbital state at a given site i. Color map indicates the strength of the
average spin-orbit, h~Li~Sii, i.e., red, yellow, green, blue, violet correspond to the growing amplitude of the above correlation
function. Small arrows at 5and10indicate quenched magnetization at the impurity by large spin-orbit coupling at the
neighboring host sites.
to thecaxis) to indicate the anisotropic spin pattern.
The out-of-plane arrow length is given by the on-site ex-
pectation value of hSz
iiwhile the in-plane arrow length
is obtained by computing the square root of the second
moment, i.e.,p
h(Sx
i)2iandp
h(Sy
i)2iof thexandyspin
components corresponding to the arrows along aandb,
respectively.
Moreover, the in-plane arrow orientation for a given
direction is determined by the sign of the correspond-
ing spin-spin correlation function assuming as a refer-ence the orientation of the impurity spin. The ellipsoid
is constructed in the same way as for the zero spin-orbit
case above, with the addition of a color map that indi-
cates the strength of the average ~Li~Si(i.e., red, yel-
low, green, blue, violet correspond with a growing ampli-
tude of the local spin-orbit correlation function). The
scale for the spin-orbit amplitude is set to be in the
interval 0 < < J host. The selected values for the
ground state evolution are given by the relation (with19
m= 1;2;:::; 10),
m=
0:04 + 0:96(m 1)
9
Jhost: (28)
The scale is set such that 1= 0:04Jhostand10=Jhost.
This range of values allows us to explore the relevant
physical regimes when moving from 3 dto 4dand 5dmate-
rials with corresponding being much smaller that Jhost,
Jhost=2 and > J host, respectively. For the per-
formed analysis the selected values of (28) are also rep-
resentative of the most interesting regimes of the ground
state as induced by the spin-orbit coupling.
Let us start with the quantum AF phases QAF c1,
QAFc2, QFSa1, QFSa2, QAFa1, and QAF a2. As one
can observe the switching on of the spin-orbit coupling
(i.e.,1in Fig. 14) leads to anisotropic spin patterns
with unequal moments for the in-plane and out-of-plane
components. From weak to strong spin-orbit coupling,
the character of the spin correlations keeps being AF be-
tween the impurity and the neighboring host sites in all
the spin directions. The main change for the spin sector
occurs for the planar components. For weak spin-orbit
coupling the in-plane spin pattern is generally AF for the
whole system in all the spatial directions (i.e., G-AF or-
der). Further increase of the spin-orbit does not modify
qualitatively the character of the spin pattern for the out-
of-plane components as long as we do not go to maximal
values ofJhostwhere localhSz
iimoments are strongly
suppressed. In this limit the dominant tendency of the
system is towards formation of the spin-orbital singlets
and the spin patterns shown in Fig. 14 are the eect of
the virtual singlet-triplet excitations [67].
Concerning the orbital sector, only for weak spin-orbit
coupling around the impurity one can still observe a rem-
iniscence of inactive orbitals as related to the orbital va-
cancy role at the impurity site in the AF phase. Such
an orbital conguration is quickly modied by increasing
the spin-orbit interaction and it evolves into a uniform
pattern with almost degenerate orbital occupations in all
the directions, and with preferential superpositions of c
and (a;b) states associated with dominating LxandLy
orbital angular components (
attened ellipsoids along the
cdirection). An exception is the QAF c2 phase with the
orbital inactive polaron that is stable up to large spin-
orbit coupling of the order of Jhost.
When considering the quantum FM congurations
QFMa1 in Fig. 14, we observe similar trends in the evo-
lution of the spin correlation functions as obtained for
the AF states. Indeed, the QFM aexhibits a tendency to
form FM chains with AF coupling for the in-plane compo-
nents at weak spin-orbit that evolve into more dominant
AF correlations in all the spatial directions within the
host. Interestingly, the spin exchange between the impu-
rity and the neighboring host sites shows a changeover
from AF to FM for the range of intermediate-to-strong
spin-orbit amplitudes.
A peculiar response to the spin-orbit coupling is ob-
tained for the QFS a1 phase, see Fig. 15, which showed
QFSa2 QFSa1
λ1
λ2
λ3λ4
λ8
λ10λ1
λ2
λ3
λ5
λ6
λ10a
ba
bFIG. 15. Evolution of the ground state congurations for
the QFSa1 and QFSa2 phases for selected increasing values
of spin-orbit coupling m, see Eq. (28). Arrows and ellip-
soids indicate the spin-orbital state at a given site i. Color
map indicates the strength of the average spin-orbit, h~Li~Sii,
i.e., red, yellow, green, blue, violet correspond to the growing
amplitude of the above local correlation function.
a frustrated spin pattern around the impurity already in
the classical regime, with FM and AF bonds. It is re-
markable that due to the close proximity with uniform
FM and the AF states, the spin-orbit interaction can lead
to a dramatic rearrangement of the spin and orbital cor-
relations for such a conguration. At weak spin-orbit
coupling (i.e., '1) the spin-pattern is C-AF and the
increased coupling ( '2)) keeps the C-AF order only20
for the in-plane components with the exception of the im-
purity site. It also modulates the spin moment distribu-
tion around the impurity along the zdirection. Further
increase of leads to complete spin polarization along
thezdirection in the host, with antiparallel orientation
with respect to the impurity spin. This pattern is guided
by the proximity to the FM phase. The in-plane compo-
nents develop a mixed FM-AF pattern with a strong xy
anisotropy most probably related to the dierent bond
exchange between the impurity and the host.
When approaching the regime of a spin-orbit coupling
that is comparable to Jhost, the out-of-plane spin compo-
nents dominate and the only out-of-plane spin polariza-
tion is observed at the impurity site. Such a behavior is
unique and occurs only in the QFS aphases. The coop-
eration between the strong spin-orbit coupling and the
frustrated host-impurity spin-orbital exchange leads to
an eective decoupling in the spin sector at the impu-
rity with a resulting maximal polarization. On the other
hand, as for the AF states, the most favorable congu-
ration for strong spin-orbit has AF in-plane spin corre-
lations. The orbital pattern for the QFS astates evolves
similarly to the AF cases with a suppression of the active-
inactive interplay around the impurity and the setting of
a uniform-like orbital conguration with unquenched an-
gular momentum on site and predominant in-plane com-
ponents. The response of the FM state is dierent in
this respect as the orbital active states around the impu-
rity are hardly aected by the spin-orbit while the host
sites far from the impurity the local spin-orbit coupling
is more pronounced.
Finally, to understand the peculiar evolution of the
spin conguration it is useful to consider the lowest order
terms in the spin-orbital exchange that couple directly
the orbital angular momentum with the spin. Taking into
account the expression of the spin-orbital exchange in the
host (26) and the expression of ~Lione can show that the
low energy terms on a bond that get more relevant in
the Hamiltonian when the spin-orbit coupling makes a
non-vanishing local angular momentum. As a result, the
corresponding expressions are:
Ha(b)
host(i;j)Jhostn
a1~Si~Sj+b1Sz
iSz
jLy(x)
iLy(x)
jo
+n
~Li~Si+~Lj~Sjo
; (29)
with positive coecients a1andb1that depend on r1and
r2(21). A denite sign for the spin exchange in the limit
of vanishing spin-orbit coupling is given by the terms
which go beyond Eq. (29). Then, if the ground state
has isotropic FM correlations (e.g. QFM a) at= 0, the
termSz
iSz
jLy(x)
iLy(x)
jwould tend to favor AF-like con-
gurations for the in-plane orbital angular components
when the spin-orbit interaction is switched on. This op-
posite tendency between the zandfx;ygcomponents is
counteracted by the local spin-orbit coupling that pre-
vents to have coexisting FM and AF spin-orbital corre-
lations. Such patterns would not allow to optimize the~Li~Siamplitudes. One way out is to reduce the zthspin
projection and to get planar AF correlations in the spin
and in the host. A similar reasoning applies to the AF
states where the negative sign of the Sz
iSz
jcorrelations
favors FO alignment of the angular momentum compo-
nents. As for the previous case, the opposite trend of in-
and out-of-plane spin-orbital components is suppressed
by the spin-orbit coupling and the in-plane FO correla-
tions for thefLx;Lygcomponents leads to FM patterns
for the in-plane spin part as well.
Summarizing, by close inspection of Figs. 14 and 15
one nds an interesting evolution of the spin patterns in
the quantum phases:
(i) For the QAF states (Fig. 14), a spin canting devel-
ops at the host sites (i.e., the relative angle is between
0 and) while the spins on impurity-host bonds are al-
ways AF. The canting in the host evolves, sometime in
an inhomogeneous way, to become reduced in the strong
spin-orbit coupling regime where ferro-like correlations
tend to dominate. In this respect, when the impurity is
coupled antiferromagnetically to the host it does not fol-
low the tendency to form spin canting.
(ii) In the QFM states (Fig. 15), at weak spin-orbit
one observes spin-canting in the host and for the host-
impurity coupling that persists only in the host whereas
the spin-orbit interaction is increasing.
C. Spin-orbit coupling versus Hund's exchange
To probe the phase diagram of the system in presence
of the spin-orbit coupling ( > 0) we solved the same
cluster ofL= 8 sites as before along three dierent cuts
in the phase diagram of Fig. 13(a) for three values of ,
i.e., small= 0:1Jhost, intermediate = 0:5Jhost, and
large=Jhost. Each cut contained ten points, the cuts
were parameterized as follows: (i) Jimp= 0:7Jhostand
0imp0:7, (ii)Jimp= 1:3Jhostand 0imp0:7,
and (iii)imp=c
imp'0:43 and 0Jimp1:5Jhost.
In Fig. 16(a) we show the representative spin-orbital
congurations obtained for = 0:5Jhostalong the rst
cut shown in Fig. 16(b). Values of impare chosen as
imp=m0:7(m 1)
9; (30)
withm= 1;:::; 10 but not all the points are shown in
Fig. 16(a) | only the ones for which the spin-orbital
conguration changes substantially.
The cut starts in the QAF c2 phase, according to the
phase diagram of Fig. 16(b), and indeed we nd a simi-
lar conguration to the one shown in Fig. 14 for QAF c2
phase at=5. Moving up in the phase diagram from
1to2we see that the conguration evolves smoothly
to the one which we have found in the QAF a1 phase at
=5(not shown in Fig. 14). The evolution of spins is
such that the out-of-plane moments are suppressed while
in-plane ones are slightly enhanced. The orbitals become
more spherical and the local spin-orbit average, h~Li~Sii,21
FIG. 16. (a) Evolution of the ground state congurations as
for increasing impand for a xed value of spin-orbit coupling
= 0:5Jhostalong a cut in the phase diagram shown in panel
(b), i.e., for Jimp= 0:7Jhostand 0imp0:7. Arrows
and ellipsoids indicate the spin-orbital state at a given site
i. Color map indicates the strength of the average spin-orbit,
h~Li~Sii, i.e., red, yellow, green, blue, violet correspond to the
growing amplitude of the above correlation function.
becomes larger and more uniform, however for the apical
sitei= 7 in the cluster [Fig. 13(b)] the trend is oppo-
site | initially large value of spin-orbit coupling drops
towards the uniform value. The points between 3and
7we skip as the evolution is smooth and the trend is
clear, however the impurity out-of-plane moment begins
to grow above 5, indicating proximity to the QFS a1
phase. For this phase at intermediate and high the
impurity moment is much larger than all the others (see
Fig. 15).
Forimp=7the orbital pattern clearly shows that
we are in the QFS a1 phase at=5which agrees with
the position of the 7point in the phase diagram, see
Fig. 16(b). On the other hand, moving to the next imp
point upward along the cut Eq. (30) we already observe
a conguration which is very typical for the QFM aphase
at intermediate (here=7shown in Fig. 14 but also
6, not shown). This indicates that the QFS a1 phase can
be still distinguished at = 0:5Jhostand its position in
the phase diagram is similar as in the = 0 case, i.e., as
an intermediate phase between the QAF a1(2) and QFM a
one.
Finally, we have found that also the two other cuts
which were not shown here, i.e., for Jimp= 1:3Jhostand
increasingimpand forimp=c
imp'0:43 and increas-
ingJimpconrm that the overall character of the phase
diagram of Fig. 13(a) is preserved at this value of spin-orbit coupling, however rstly, the transitions between
the phases are smooth and secondly, the subtle dier-
ences between the two QFS a, QAFaand QAFcphases
are no longer present. This also refers to the smaller
value of, i.e.,= 0:1Jhost, but already for =Jhost
the out-of-plane moments are so strongly suppressed (ex-
cept for the impurity moment in the QFS a1 phase) and
the orbital polarization is so weak (i.e., almost spheri-
cal ellipsoids) that typically the only distinction between
the phases can be made by looking at the in-plane spin
correlations and the average spin-orbit, h~Li~Sii. In this
limit we conclude that the phase diagram is (partially)
melted by large spin-orbit coupling but for lower values
ofit is still valid.
VI. SUMMARY AND CONCLUSIONS
We have derived the spin-orbital superexchange model
for 3d3impurities replacing 4 d4(or 3d2) ions in the 4 d
(3d) host in the regime of Mott insulating phase. Al-
though the impurity has no orbital degree of freedom, we
have shown that it contributes to the spin-orbital physics
and in
uences strongly the orbital order. In fact, it tends
to project out the inactive orbitals at the impurity-host
bonds to maximize the energy gain from virtual charge
uctuations. In this case the interaction along the su-
perexchange bond can be either antiferromagnetic or fer-
romagnetic, depending on the ratio of Hund's exchange
coupling at impurity ( JH
1) and host ( JH
2) ions and on
the mismatch between the 3 dand 4datomic energies,
modied by the dierence in Hubbard U's and Hund's
exchangeJH's at both atoms. This ratio, denoted imp
(14), replaces here the conventional parameter =JH=U
often found in the spin-orbital superexchange models of
undoped compounds (e.g., in the Kugel-Khomskii model
for KCuF 3[14]) where it quanties the proximity to fer-
romagnetism. On the other hand, if the overall coupling
between the host and impurity is weak in the sense of
the total superexchange, Jimp, with respect to the host
value,Jhost, the orbitals being next to the impurity may
be forced to stay inactive which modies the magnetic
properties | in such cases the impurity-host bond is al-
ways antiferromagnetic.
As we have seen in the case of a single impurity, the
above two mechanisms can have a nontrivial eect on
the host, especially if the host itself is characterized by
frustrated interactions, as it happens in the parameter
regime where the C-AF phase is stable. For this rea-
son we have focused mostly on the latter phase of the
host and we have presented the phase diagrams of a sin-
gle impurity conguration in the case when the impu-
rity is doped on the sublattice where the orbitals form
a checkerboard pattern with alternating candaorbitals
occupied by doublons. The diagram for the c-sublattice
doping shows that in some sense the impurity is never
weak, because even for a very small value of Jimp=Jhost
it can release the host's frustration around the impurity22
site acting as an orbital vacancy. On the other hand, for
thea-sublattice doping when the impurity-host coupling
is weak, i.e., either Jimp=Jhostis weak orimpis close to
c
imp, we have identied an interesting quantum mech-
anism releasing frustration of the impurity spin (that
cannot be avoided in the purely classical approach). It
turned out that in such situations the orbital
ips in the
host make the impurity spin polarize in such a way that
theC-AF order of the host is completely restored.
The cases of the periodic doping studied in this pa-
per show that the host's order can be completely altered
already for rather low doping ofx= 1=8, even if the
Jimp=Jhostis small. In this case we can stabilize a ferri-
magnetic type of phase with a four-site unit cell having
magnetizationhSz
ii= 3=2, reduced further by quantum
uctuations. We have established that the only param-
eter range where the host's order remains unchanged is
whenimpis close toc
impandJimp=Jhost&1. The latter
value is very surprising as it means that the impurity-host
coupling must be large enough to keep the host's order
unchanged | this is another manifestation of the orbital
vacancy mechanism that we have already observed for
a single impurity. Also in this case the impurity spins
are xed with the help of orbital
ips in the host that
lift the degeneracy which arises in the classical approach.
We would like to point out that the quantum mechanism
that lifts the ground state degeneracy mentioned above
and the role of quantum
uctuations are of particular
interest for the periodically doped checkerboard systems
withx= 1=2 doping which is a challenging problem for
future research.
From the point of view of generic, i.e., non-periodic
doping, the most representative cases are those of a
doping which is incommensurate with the two-sublattice
spin-orbital pattern. To uncover the generic rules in such
cases, we have studied periodic x= 1=5 andx= 1=9 dop-
ing. One nds that when the period of the impurity posi-
tions does not match the period of 2 for both the spin and
orbital order of the host, interesting novel types of order
emerge. In such cases the elementary cell must be dou-
bled in both lattice directions which clearly gives a chance
of realizing more phases than in the case of commensu-
rate doping. Our results show that indeed, the number of
phases increases from 4 to 7 and the host's order is altered
in each of them. Quite surprisingly, the overall character
of the phase diagram remained unchanged with respect
to the one for x= 1=8 doping and, if we ignore the dif-
ferences in conguration, it seems that only some of the
phases got divided into two versions diering either by
the spin bond's polarizations around impurities (phases
aroundc
imp), or by the character of the orbitals around
the impurities (phases with inactive orbitals in the limit
of small enough product impJimp, versus phases with ac-
tive orbitals in the opposite limit). Orbital polarization
in this latter region resembles orbital polarons in doped
manganites [42, 43] | also here such states are stabilized
by the double exchange [46].
A closer inspection of underlying phases reveals how-ever a very interesting degeneracy of the impurity spins
atx= 1=5 that arises again from the classical approach
but this time it cannot be released by short-range orbital
ips. This happens because the host's order is already
so strongly altered that it is no longer anisotropic (as it
was the case of the C-AF phase) and there is no way
to restore the orbital anisotropy around the impurities
that could lead to spin-bonds imbalance and polarize the
spin. In the case of lower x= 1=9 doping such an eect
is absent and the impurity spins are always polarized, as
it happens for x= 1=8. It shows that this is rather a
peculiarity of the x= 1=5 periodic doping.
Indeed, one can easily notice that for x= 1=5 every
atom of the host is a nearest neighbor of some impurity.
In contrast, for x= 1=8 we can nd three host's atoms
per unit cell which do not neighbor any impurity and for
x= 1=9 there are sixteen of them. For this reason the
impurity eects are amplied for x= 1=5 which is not
unexpected although one may nd somewhat surprising
that the ground state diagrams for the lowest and the
highest doping considered here are very similar. This
suggests that the cooperative eects of multiple impuri-
ties are indeed not very strong in the low-doping regime,
so the diagram obtained for x= 1=9 can be regarded as
generic for the dilute doping regime with uniform spatial
prole.
For the representative case of x= 1=8 doping, we have
presented the consequences of quantum eects beyond
the classical approach. Spin
uctuations are rather weak
for the considered case of large S= 1 andS= 3=2 spins,
and we have shown that orbital
uctuations on superex-
change bonds are more important. They are strongest in
the regime of antiferromagnetic impurity-host coupling
(which suggests importance of entangled states [34]) and
enhance the tendency towards frustrated impurity spin
congurations but do not destroy other generic trends
observed when the parameters impandJimp=Jhostin-
crease.
Increasing spin-orbit coupling leads to qualitative
changes in the spin-orbital order. When Hund's ex-
change is small at the impurity sites, the antiferromag-
netic bonds around it have reduced values of spin-orbit
coupling term, but the magnetic moments reorient and
survive in the ( a;b) planes, with some similarity to the
phenomena occurring in the perovskite vanadates [57].
This quenches the magnetic moments at 3 dimpurities
and leads to almost uniform orbital occupancies at the
host sites. In contrast, frustration of impurity spins is re-
moved and the impurity magnetization along the caxis
survives for large spin-orbit coupling.
We would like to emphasize that the orbital dilution
considered here in
uences directly the orbital degrees of
freedom in the host around the impurities. The synthesis
of hybrid compounds having both 3 dand 4dtransition
metal ions will likely open a novel route for unconven-
tional eects in complex materials. There are several
reasons for expecting new scenarios in mixed 3 d 4d
spin-orbital-lattice materials, and we pointed out only23
some of them. On the experimental side, the changes
of local order could be captured using inelastic neutron
scattering or resonant inelastic x-ray scattering (RIXS).
In fact, using RIXS can also bring an additional advan-
tage: RIXS, besides being a perfect probe of both spin
and orbital excitations, can also (indirectly) detect the
nature of orbital ground state (supposedly also including
the nature of impurities in the crystal) [82]. Unfortu-
nately, there are no such experiments yet but we believe
that they will be available soon.
Short range order around impurities could be inves-
tigated by the excitation spectra at the resonant edges
of the substituting atoms. Taking them both at nite
energy and momentum can dive insights into the nature
of the short range order around the impurity and then
unveil information of the order within the host as well.
Even if there are no elastic superlattice extra peaks one
can expect that the spin-orbital correlations will emerge
in the integrated RIXS spectra providing information of
the impurity-host coupling and of the short range order
around the impurity. Even more interesting is the case
where the substituting atom forms a periodic array with
small deviation from the perfect superlattice when one
expects the emergence of extra elastic peaks which will
clearly indicate the spin-orbital reconstruction. In our
case an active orbital diluted site cannot participate co-
herently in the host spin-orbital order but rather may
to restructure the host ordering [83]. At dilute impurity
concentration we may expect broad peaks emerging at
nite momenta in the Brillouin zone, indicating the for-
mation of coherent islands with short range order around
impurities.
We also note that local susceptibility can be suit-
ably measured by making use of resonant spectroscopies
(e.g. nuclear magnetic resonance (NMR), electron spin
resonance (ESR), nuclear quadrupole resonance (NQR),
muon spin resonance ( SR), etcetera ) that exploit the
dierent magnetic or electric character of the atomic nu-
clei for the impurity and the host in the hybrid system.
Finally, the random implantation of the muons in the
sample can provide information of the relaxation time
in dierent domains with unequal dopant concentration
which may be nonuniform. For the given problem the
dierences in the resonant response can give relevant in-
formation about the distribution of the local elds, the
occurrence of local order and provide access to the dy-
namical response within doped domains. The use of local
spectroscopic resonance methods has been widely demon-
strated to be successful when probing the nature and the
evolution of the ground state in the presence of spin va-
cancies both for ordered and disordered magnetic cong-
urations [84{87].
In summary, this study highlights the role of spin de-
fects which lead to orbital dilution in spin-orbital sys-
tems. Using an example of 3 d3impurities in a 4 d4(or
3d2) host we have shown that impurities change radi-
cally the spin-orbital order around them, independently
of the parameter regime. As a general feature we havefound that doped 3 d3ions within the host with spin-
orbital order have frustrated spins and polarize the or-
bitals of the host when the impurity-host exchange as
well as Hund's exchange at the impurity are both suf-
ciently large. This remarkable trend is independent of
doping and is expected to lead to global changes of spin-
orbital order in doped materials. While the latter eect
is robust, we argue that the long-range spin
uctuations
resulting from the translational invariance of the system
will likely prevent the ground state from being macro-
scopically degenerate, so if the impurity spins in one unit
cell happens to choose its polarization then the others will
follow. On the contrary, in the regime of weak Hund's
exchange 3d3ions act not only as spin defects which or-
der antiferromagnetically with respect to their neighbors,
but also induce doublons in inactive orbitals.
Finally, we remark that this behavior with switching
between inactive and active orbitals by an orbitally neu-
tral impurity may lead to multiple interesting phenomena
at macroscopic doping when global modications of the
spin-orbital order are expected to occur. Most of the re-
sults were obtained in the classical approximation but we
have shown that modications due to spin-orbit coupling
do not change the main conclusion. We note that this
generic treatment and the general questions addressed
here, such as the release of frustration for competing
spin structures due to periodic impurities, are relevant
to double perovskites [88]. While the local orbital polar-
ization should be similar, it is challenging to investigate
disordered impurities, both theoretically and in experi-
ment, to nd out whether their in
uence on the global
spin-orbital order in the host is equally strong.
ACKNOWLEDGMENTS
We thank Maria Daghofer and Krzysztof Wohlfeld for
insightful discussions. W. B. and A. M. O. kindly ac-
knowledge support by the Polish National Science Cen-
ter (NCN) under Project No. 2012/04/A/ST3/00331.
W. B. was also supported by the Foundation for Pol-
ish Science (FNP) within the START program. M. C.
acknowledges funding from the EU | FP7/2007-2013
under Grant Agreement No. 264098 | MAMA.
Appendix A: Derivation of 3d 4dsuperexchange
Here we present the details of the derivation of the low
energy spin-orbital Hamiltonian for the 3 d3 4d4bonds
around the impurity at site i.H3d 4d(i), which follows
from the perfurbation theory, as given in Eq. (10). Here
we consider a single 3 d3 4d4bondhiji. Two contri-
butions to the eective Hamiltonian follow from charge
excitations: (i)H(
)
J;43(i;j) due tod3
id4
jd4
id3
j, and (ii)
H(
)
J;25(i;j) due tod3
id4
jd2
id5
j. Therefore the low energy24
Hamiltonian is,
H(
)
J(i;j) =H(
)
J;43(i;j) +H(
)
J;25(i;j): (A1)
Consider rst the processes which conserve the num-
ber of doubly occupied orbitals, d3
id4
jd4
id3
j. Then by
means of spin and orbital projectors, it is possible to ex-
pressH(
)
J;43(i;j) fori= 1 andj= 2 as
H(
)
J;43(1;2) =
~S1~S2t2
184
7
+ 3JH
2 3
+ 5JH
2
+D(
)
2
~S1~S2t2
184
1
+ 3JH
2+3
+ 5JH
2
+
D(
)
2 1t2
128
+1
+ 3JH
2 3
+ 5JH
2
;(A2)
with the excitation energy dened in Eq. (11). The
resulting eective 3 d 4dexchange in Eq. (A2) consists
of three terms: (i) The rst one does not depend on the
orbital conguration of the 4 datom and it can be FM or
AF depending on the values and the Hund's exchange
on the 3dion. In particular, if is the largest or the
smallest energy scale, the coupling will be either AF or
FM, respectively. (ii) The second term has an explicit
dependence on the occupation of the doublon on the 4 d
atom via the projecting operator D(
)
2. This implies that
a magnetic exchange is possible only if the doublon occu-
pies the inactive orbital for a bond along a given direction
. Unlike in the rst term, the sign of this interaction
is always positive favoring an AF conguration at any
strength of and JH
1. (iii) Finally, the last term de-
scribes the eective processes which do not depend on
the spin states on the 3 dand 4datoms. This contri-
bution is of pure orbital nature, as it originates from the
hopping between 3 dand 4datoms without aecting their
spin conguration, and for this reason favors the occupa-
tion of active t2gorbitals along the bond by the doublon.
Within the same scheme, we have derived the ef-
fective spin-orbital exchange that originates from the
charge transfer processes of the type 3 d3
14d4
23d2
i4d5
j,
H(
)
J;25(1;2). The eective low-energy contribution to the
Hamiltonian for i= 1 andj= 2 reads
H(
)
J;25(1;2) =t2
U1+U2
+ 3JH
2 2JH
1
1
3D(
)
2
~S1~S2
+1
3
~S1~S2
1
2
D(
)
2+ 1
:(A3)
By inspection of the spin structure involved in the ele-
mental processes that generate H(
)
J;25(1;2), one can note
that it is always AF independently of the orbital con-
guration on the 4 datom exhibiting with a larger spin-
exchange and an orbital energy gain if the doublon is
occupying the inactive orbital along a given bond. We
have veried that the amplitude of the exchange termsinH(
)
J;25(1;2) is much smaller than the ones which enter
inH(
)
J;43(1;2) which justies that one may simplify Eq.
(A1) fori= 1 andj= 2 to
H(
)
J(1;2)'H(
)
J;43(1;2); (A4)
and neglectH(
)
J;25(1;2) terms altogether. This approxi-
mation is used in Sec. II.
Appendix B: Orbital operators in the L-basis
The starting point to express the orbital operators ap-
pearing in the spin-orbital superexchange model (17) is
the relation between quenched jaii,jbii, andjciiorbitals
at siteiand the eigenvectors j1ii,j0ii, andj 1iiof the
angular momentum operator Lz
i. These are known to be
jaii=1p
2(j1ii+j 1ii);
jbii= ip
2(j1ii j 1ii);
jcii=j0ii: (B1)
From this we can immediately get the occupation number
operators for the doublon,
D(a)
i=ay
iai=jaiihaji= 1 (Lx
i)2;
D(b)
i=by
ibi=jbiihbji= 1 (Ly
i)2;
D(c)
i=cy
ici=jciihcji= 1 (Lz
i)2; (B2)
and the relatedfn(
)
igoperators,
n(a)
i=by
ibi+cy
ici= (Lx
i)2;
n(b)
i=cy
ici+ay
iai= (Ly
i)2;
n(c)
i=ay
iai+by
ibi= (Lz
i)2: (B3)
The doublon hopping operators have a slightly dierent
structure that re
ects their noncommutivity, i.e.,
ay
ibi=jaiihbji=iLy
iLx
i;
by
ici=jbiihcji=iLz
iLy
i;
cy
iai=jciihaji=iLx
iLz
i: (B4)
These relations are sucient to write the superexchange
Hamiltonian for the host-host and impurity-host bonds
in thefLx
i;Ly
i;Lz
igoperator basis for the orbital part.
However, in practice it is more convenient to work with
real operators
L+
i;L
i;Lz
i
rather than with the origi-
nal ones,fLx
i;Ly
i;Lz
ig. Thus we write the nal relations
which we used for the numerical calculations in terms of
these operators,
D(a)
i= 1
4h
L+
i2+
L
i2i
+1
2(Lz
i)2;
D(b)
i=1
4h
L+
i2+
L
i2i
+1
2(Lz
i)2;25
D(c)
i= 1 (Lz
i)2; (B5)
for the doublon occupation numbers and going directly
to the orbital ~ ioperators we nd that,
+(a)
i =1
2
L
i L+
i
Lz
i;
+(b)
i= i
2Lz
i
L+
i+L
i
;
+(c)
i=i
4h
L+
i2
L
i2i
i
2Lz
i; (B6)
for the o-diagonal part and
z(a)
i=1
8h
L+
i2+
L
i2i
+3
4(Lz
i)2 1
2;
z(b)
i=1
8h
L+
i2+
L
i2i
3
4(Lz
i)2+1
2;z(c)
i= 1
4h
L+
i2+
L
i2i
; (B7)
for the diagonal one. Note that the complex phase in
+(b)
i and+(c)
i is irrelevant and can be omitted here as
+(
)
i is always accompanied by (
)
j on a neighboring
site. This is a consequence of the cubic symmetry in the
orbital part of the superexchange Hamiltonian and it can
be altered by a presence of a distortion, e.g., octahedral
rotation. For completeness we also give the backward
relation between angular momentum components, fL
ig
with=x;y;z , and the orbital operators f(
)
ig; these
are:
Lx
i= 2x(a)
i;
Ly
i= 2x(b)
i;
Lz
i= 2y(c)
i: (B8)
[1] M. Imada, A. Fujimori, and Y. Tokura, Metal-Insulator
Transitions , Rev. Mod. Phys. 70, 1039 (1998).
[2] E. Berg, E. Fradkin, S. A. Kivelson, and J. M. Tran-
quada, Striped Superconductors: how Spin, Charge and
Superconducting Orders Intertwine in the Cuprates , New
J. Phys. 11, 115004 (2011).
[3] P. A. Lee, N. Nagaosa, and X.-G. Wen, Doping a Mott
Insulator: Physics of High-Temperature Superconductiv-
ity, Rev. Mod. Phys. 78, 17 (2006).
[4] M. Vojta, Lattice Symmetry Breaking in Cuprate Super-
conductors: Stripes, Nematics, and Superconductivity ,
Adv. Phys. 58, 699 (2009).
[5] K. Yamada, C. H. Lee, K. Kurahashi, J. Wada, S. Waki-
moto, S. Ueki, H. Kimura, Y. Endoh, S. Hosoya, G. Shi-
rane, R. J. Birgeneau, M. Greven, M. A. Kastner, and
Y. J. Kim, Doping Dependence of the Spatially Modulated
Dynamical Spin Correlations and the Superconducting-
Transition Temperature in La2 xSrxCuO 4, Phys. Rev. B
57, 6165 (1998).
[6] M. Fleck, A. I. Lichtenstein, and A. M. Ole s, Spectral
Properties and Pseudogap in the Stripe Phases of Cuprate
Superconductors , Phys. Rev. B 64, 134528 (2001).
[7] P. Wr obel and A. M. Ole s, Ferro-Orbitally Ordered
Stripes in Systems with Alternating Orbital Order , Phys.
Rev. Lett. 104, 206401 (2010).
[8] Q. Li, K. E. Gray, S. B. Wilkins, M. Garcia Fernan-
dez, S. Rosenkranz, H. Zheng, and J.F. Mitchell, Predic-
tion and Experimental Evidence for Thermodynamically
Stable Charged Orbital Domain Walls , Phys. Rev. X 4,
031028 (2014).
[9] J. Zaanen and A. M. Ole s, Carriers Binding to Excitons:
Crystal-Field Excitations in Doped Mott-Hubbard Insula-
tors, Phys. Rev. B 48, 7197 (1993).
[10] Y. Tokura and N. Nagaosa, Orbital Physics in
Transition-Metal Oxides , Science 288, 462 (2000).
[11] K. I. Kugel and D. I. Khomskii, The Jahn-Teller Ef-
fect and Magnetism: Transition Metal Compounds , Sov.
Phys. Usp. 25, 231 (1982).[12] A. M. Ole s, G. Khaliullin, P. Horsch, and L. F. Feiner,
Fingerprints of Spin-Orbital Physics in Cubic Mott In-
sulators: Magnetic Exchange Interactions and Optical
Spectral Weights , Phys. Rev. B 72, 214431 (2005).
[13] K. Wohlfeld, M. Daghofer, S. Nishimoto, G. Khaliullin,
and J. van den Brink. Intrinsic Coupling of Orbital Ex-
citations to Spin Fluctuations in Mott Insulators , Phys.
Rev. Lett. 107, 147201 (2011).
[14] L. F. Feiner, A. M. Ole s, and J. Zaanen, Quantum Melt-
ing of Magnetic Order due to Orbital Fluctuations , Phys.
Rev. Lett. 78, 2799 (1997).
[15] G. Khaliullin and S. Maekawa, Orbital Liquid in Three-
Dimensional Mott Insulator: LaTiO 3, Phys. Rev. Lett.
85, 3950 (2000).
[16] G. Khaliullin, Orbital Order and Fluctuations in Mott
Insulators , Prog. Theor. Phys. Suppl. 160, 155 (2005).
[17] Leon Balents, Spin Liquids in Frustrated Magnets , Na-
ture (London) 464, 199 (2010).
[18] L. F. Feiner and A. M. Ole s, Orbital Liquid in Ferro-
magnetic Manganites: The Orbital Hubbard Model for eg
Electrons , Phys. Rev. B 71, 144422 (2005).
[19] S. Nakatsuji, K. Kuga, K. Kimura, R. Satake,
N. Katayama, E. Nishibori, H. Sawa, R. Ishii, M. Hagi-
wara, F. Bridges, T. U. Ito, W. Higemoto, Y. Karaki, M.
Halim, A. A. Nugroho, J. A. Rodriguez-Rivera, M. A.
Green, and C. Broholm, Spin-Orbital Short-Range Order
on a Honeycomb-Based Lattice , Science 336, 559 (2012).
[20] J. A. Quilliam, F. Bert, E. Kermarrec, C. Payen,
C. Guillot-Deudon, P. Bonville, C. Baines, H. Luetkens,
and P. Mendels, Singlet Ground State of the Quantum
Antiferromagnet Ba3CuSb 2O9, Phys. Rev. Lett. 109,
117203 (2012).
[21] B. Normand and A. M. Ole s, Frustration and Entangle-
ment in the t2gSpin-Orbital Model on a Triangular Lat-
tice: Valence-bond and Generalized Liquid States , Phys.
Rev. B 78, 094427 (2008).
[22] P. Corboz, M. Lajk o, A. M. La uchli, K. Penc, and
F. Mila, Spin-Orbital Quantum Liquid on the Honeycomb
Lattice , Phys. Rev. X 2, 041013 (2012).26
[23] J. Nasu and S. Ishihara, Dynamical Jahn-Teller Eect in
a Spin-Orbital Coupled System , Phys. Rev. B 88, 094408
(2012).
[24] A. Smerald and F. Mila, Exploring the Spin-Orbital
Ground State of Ba3CuSb 2O9, Phys. Rev. B 90, 094422
(2014).
[25] V. Fritsch, J. Hemberger, N. B uttgen, E.-W. Scheidt, H.-
A. Krug von Nidda, A. Loidl, and V. Tsurkan, Spin and
Orbital Frustration in MnSc 2S2andFeSc 2S4, Phys. Rev.
Lett. 92, 116401 (2004).
[26] G. Chen, L. Balents, and A. P. Schnyder, Spin-Orbital
Singlet and Quantum Critical Point on the Diamond Lat-
tice: FeSc 2S4, Phys. Rev. Lett. 102, 096406 (2009).
[27] L. Mittelst adt, M. Schmidt, Z. Wang, F. Mayr,
V. Tsurkan, P. Lunkenheimer, D. Ish, L. Balents, J.
Deisenhofer, and A. Loidl, Spin-Orbiton and Quantum
Criticality in FeSc 2S4, Phys. Rev. B 91, in press (2015).
[28] F. Vernay, K. Penc, P. Fazekas, and F. Mila, Orbital De-
generacy as a Source of Frustration in LiNiO 2, Phys.
Rev. B 70, 014428 (2004).
[29] W. Brzezicki, J. Dziarmaga, and A. M. Ole s, Non-
collinear Magnetic Order Stabilized by Entangled Spin-
Orbital Fluctuations , Phys. Rev. Lett. 109, 237201
(2012).
[30] W. Brzezicki, J. Dziarmaga, and A. M. Ole s, Exotic
Spin Orders Driven by Orbital Fluctuations in the Kugel-
Khomskii Model , Phys. Rev. B 87, 064407 (2013).
[31] P. Czarnik and J. Dziarmaga, Striped Critical Spin Liquid
in a Spin-Orbital Entangled RVB State in a Projected
Entangled-Pair State Representation , Phys. Rev. B 91,
045101 (2015).
[32] F. Reynaud, D. Mertz, F. Celestini, J. Debierre,
A. M. Ghorayeb, P. Simon, A. Stepanov, J. Voiron, and
C. Delmas, Orbital Frustration at the Origin of the Mag-
netic Behavior in LiNiO 2, Phys. Rev. Lett. 86, 3638
(2001).
[33] A. Reitsma, L. F. Feiner, and A. M. Ole s, Orbital and
Spin Physics in LiNiO 2and NaNiO 2, New J. Phys. 7,
121 (2005).
[34] A. M. Ole s, Fingerprints of Spin-Orbital Entanglement
in Transition Metal Oxides , J. Phys.: Condens. Matter
24, 313201 (2012).
[35] E. Dagotto, T. Hotta, and A. Moreo, Colossal Magne-
toresistant Materials: The Key Role of Phase Separation ,
Phys. Rep. 344, 1 (2001).
[36] E. Dagotto, Open Questions in CMR Manganites, Rel-
evance of Clustered States and Analogies with Other
Compounds Including the Cuprates , New J. Phys. 7, 67
(2005).
[37] Y. Tokura, Critical Features of Colossal Magnetoresistive
Manganites , Rep. Prog. Phys. 69, 797 (2006).
[38] L. F. Feiner and A. M. Ole s, Electronic Origin of Mag-
netic and Orbital Ordering in Insulating LaMnO 3, Phys.
Rev. B 59, 3295 (1999).
[39] R. Kilian and G. Khaliullin, Orbital Polarons in the
Metal-Insulator Transition of Manganites , Phys. Rev. B
60, 13458 (1999).
[40] M. Cuoco, C. Noce, and A. M. Ole s, Origin of the Optical
Gap in Half-Doped Manganites , Phys. Rev. B 66, 094427
(2002).
[41] A. Weisse and H. Fehske, Microscopic Modelling of Doped
Manganites , New J. Phys. 6, 158 (2004).
[42] M. Daghofer, A. M. Ole s, and W. von der Linden, Orbital
Polarons versus Itinerant egElectrons in Doped Mangan-ites, Phys. Rev. B 70, 184430 (2004).
[43] J. Geck, P. Wochner, S. Kiele, R. Klingeler, P. Reutler,
A. Revcolevschi, and B. B uchner, Orbital Polaron Lattice
Formation in Lightly Doped La1 xSrxMnO 3, Phys. Rev.
Lett. 95, 236401 (2005).
[44] A. M. Ole s and G. Khaliullin, Dimensional Crossover and
the Magnetic Transition in Electron Doped Manganites ,
Phys. Rev. B 84, 214414 (2011).
[45] Y. M. Sheu, S. A. Trugman, L. Yan, J. Qi, Q. X. Jia,
A. J. Taylor, and R. P. Prasankumar, Polaronic Trans-
port Induced by Competing Interfacial Magnetic Order in
aLa0:7Ca0:3MnO 3/BiFeO 3Heterostructure , Phys. Rev.
X4, 021001 (2014).
[46] P.-G. de Gennes, Eects of Double Exchange in Magnetic
Crystals , Phys. Rev. 118, 141 (1960).
[47] G. Khaliullin and R. Kilian, Theory of Anomalous
Magnon Softening in Ferromagnetic Manganites , Phys.
Rev. B 61, 3494 (2000).
[48] A. M. Ole s and L. F. Feiner, Why Spin Excitations in
Metallic Ferromagnetic Manganites are Isotropic , Phys.
Rev. B 65, 052414 (2002).
[49] H. Sakai, S. Ishiwata, D. Okuyama, A. Nakao, H. Nakao,
Y. Murakami, Y. Taguchi, and Y. Tokura, Electron Dop-
ing in the Cubic Perovskite SrMnO 3:Isotropic Metal ver-
sus Chainlike Ordering of Jahn-Teller Polarons , Phys.
Rev. B 82, 180409 (2010).
[50] G. Khaliullin, P. Horsch, and A. M. Ole s, Spin Order due
to Orbital Fluctuations: Cubic Vanadates , Phys. Rev.
Lett. 86, 3879 (2001).
[51] G. Khaliullin, P. Horsch, and A. M. Ole s, Theory of Op-
tical Spectral Weights in Mott Insulators with Orbital De-
grees of Freedom , Phys. Rev. B 70, 195103 (2004).
[52] P. Horsch, A. M. Ole s, L. F. Feiner, and G. Khaliullin,
Evolution of Spin-Orbital-Lattice Coupling in the RVO3
Perovskites , Phys. Rev. Lett. 100, 167205 (2008).
[53] C. Ulrich, G. Khaliullin, J. Sirker, M. Reehuis, M. Ohl,
S. Miyasaka, Y. Tokura, and B. Keimer, Magnetic Neu-
tron Scattering Study of YVO 3:Evidence for an Orbital
Peierls State , Phys. Rev. Lett. 91, 257202 (2003).
[54] J.-S. Zhou, J. B. Goodenough, J.-Q. Yan, and
Y. Ren, Superexchange Interaction in Orbitally Fluctu-
ating RVO 3, Phys. Rev. Lett. 99, 156401 (2007).
[55] M. Reehuis, C. Ulrich, K. Proke s, S. Mat'a s, J. Fu-
jioka, S. Miyasaka, Y. Tokura, and B. Keimer, Structural
and Magnetic Phase Transitions of the Orthovanadates
RVO3(R=Dy,Ho,Er) as seen via Neutron Diraction ,
Phys. Rev. B 83, 064404 (2011).
[56] J. Fujioka, T. Yasue, S. Miyasaka, Y. Yamasaki,
T. Arima, H. Sagayama, T. Inami, K. Ishii, and
Y. Tokura, Critical Competition between two Distinct
Orbital-Spin Ordered States in Perovskite Vanadates ,
Phys. Rev. B 82, 144425 (2010).
[57] P. Horsch, G. Khaliullin, and A. M. Ole s, Dimeriza-
tion versus Orbital-Moment Ordering in a Mott Insulator
YVO 3, Phys. Rev. Lett. 91, 257203 (2003).
[58] J. Fujioka, S. Miyasaka, and Y. Tokura, Doping Vari-
ation of Anisotropic Charge and Orbital Dynamics in
Y1 xCaxVO3:Comparison with La1 xSrxVO3, Phys.
Rev. B 77, 144402 (2008).
[59] P. Horsch and A. M. Ole s, Defect States and Spin-Orbital
Physics in Doped Vanadates Y1 xCaxVO3, Phys. Rev. B
84, 064429 (2011).
[60] A. Avella, P. Horsch, and A. M. Ole s, Defect States and
Excitations in a Mott Insulator with Orbital Degrees of27
Freedom: Mott-Hubbard Gap versus Optical and Trans-
port Gaps in Doped Systems , Phys. Rev. B 87, 045132
(2013).
[61] T. Hotta, Orbital Ordering Phenomena in d- andf-
Electron Systems , Rep. Prog. Phys. 69, 2061 (2006).
[62] T. Mizokawa, L. H. Tjeng, G. A. Sawatzky, G. Ghir-
inghelli, O. Tjernberg, N. B. Brookes, H. Fukazawa,
S. Nakatsuji, and Y. Maeno, Spin-Orbit Coupling in the
Mott Insulator Ca2RuO 4, Phys. Rev. Lett. 87, 077202
(2001).
[63] J. S. Lee, Y. S. Lee, T. W. Noh, S.-J. Oh, J. Yu, S. Nakat-
suji, H. Fukazawa, and Y. Maeno, Electron and Orbital
Correlations in Ca2 xSrxRuO 4Probed by Optical Spec-
troscopy , Phys. Rev. Lett. 89, 257402 (2002).
[64] A. Koga, N. Kawakami, T. M. Rice, and M. Sigrist,
Orbital-Selective Mott Transitions in the Degenerate
Hubbard Model , Phys. Rev. Lett. 92, 216402 (2004).
[65] Z. Fang, K. Terakura, and N. Nagaosa, Orbital Physics
in Ruthenates: First-Principles Study , New J. Phys. 7,
66 (2005).
[66] T. Sugimoto, D. Ootsuki, and T. Mizokawa, Impact of
Local Lattice Disorder on Spin and Orbital Orders in
Ca2 xSrxRuO 4, Phys. Rev. B 84, 064429 (2011).
[67] G. Khaliullin, Excitonic Magnetism in Van Vleck-type d4
Mott Insulators , Phys. Rev. Lett. 111, 197201 (2013).
[68] A. Akbari and G. Khaliullin, Magnetic Excitations in a
Spin-Orbit-Coupled d4Mott Insulator on the Square Lat-
tice, Phys. Rev. B 90, 035137 (2014).
[69] C. N. Veenstra, Z.-H. Zhu, M. Raichle, B. M. Lud-
brook, A. Nicolaou, B. Slomski, G. Landolt, S. Kittaka,
Y. Maeno, J. H. Dil, I. S. Elmov, M. W. Haverkort,
and A. Damascelli, Spin-Orbital Entanglement and the
Breakdown of Singlets and Triplets in Sr2RuO 4Revealed
by Spin- and Angle-Resolved Photoemission Spectroscopy ,
Phys. Rev. Lett. 112, 127002 (2014).
[70] T. Tanaka and S. Ishihara, Dilution Eects in Two-
Dimensional Quantum Orbital Systems , Phys. Rev. Lett.
98, 256402 (2007).
[71] T. F. Qi, O. B. Korneta, S. Parkin, L. E. De Long,
P. Schlottmann, and G. Cao, Negative Volume Ther-
mal Expansion via Orbital and Magnetic Orders in
Ca2Ru1 xCrxO4(0< x < 0:13), Phys. Rev. Lett. 105,
177203 (2010).
[72] G. Cao, S. Chikara, X. N. Lin, E. Elhami, V. Durairaj,
and P. Schlottmann, Itinerant Ferromagnetism to Insu-
lating Antiferromagnetism: A Magnetic and Transport
Study of Single Crystal SrRu 1 xMnxO3(0x<0:60),
Phys. Rev. B 71, 035104 (2005).
[73] M. A. Hossain, B. Bohnenbuck, Y. D. Chuang,
M. W. Haverkort, I. S. Elmov, A. Tanaka, A. G. Cruz
Gonzalez, Z. Hu, H.-J. Lin, C. T. Chen, R. Math-
ieu, Y. Tokura, Y. Yoshida, L. H. Tjeng, Z. Hus-
sain, B. Keimer, G. A. Sawatzky, and A. Damascelli,
Mott versus Slater-type Metal-Insulator Iransition in Mn-
Substituted Sr3Ru2O7, Phys. Rev. B 86, 041102(R)
(2012).
[74] M. A. Hossain, Z. Hu, M. W. Haverkort, T. Bur-
nus, C. F. Chang, S. Klein, J. D. Denlinger, H.-
J. Lin, C. T. Chen, R. Mathieu, Y. Kaneko, Y. Tokura,
S. Satow, Y. Yoshida, H. Takagi, A. Tanaka, I. S. Elmov,
G. A. Sawatzky, L. H. Tjeng, and A. Damascelli, Crystal-
Field Level Inversion in Lightly Mn-Doped Sr3Ru2O7,Phys. Rev. Lett. 101, 016404 (2008).
[75] D. Mesa, F. Ye, S. Chi, J. A. Fernandez-Baca, W. Tian,
B. Hu, R. Jin, E. W. Plummer, and J. Zhang, Single-
Bilayer E-type Antiferromagnetism in Mn-Substituted
Sr3Ru2O7:Neutron Scattering Study , Phys. Rev. B 85,
180410(R) (2012).
[76] S. A. J. Kimber, J. A. Rodgers, H. Wu, C. A. Murray,
D. N. Argyriou, A. N. Fitch, D. I. Khomskii, and J. P.
Atteld, Metal-Insulator Transition and Orbital Order in
PbRuO 3, Phys. Rev. Lett. 102, 046409 (2009).
[77] J. F. Annett, G. Litak, B. L. Gy ory, and
K. I. Wysoki nski, Spin-Orbit Coupling and Symmetry of
the Order Parameter in Strontium Ruthenate , Phys. Rev.
B73, 134501 (2006).
[78] A. B. Harris, T. Yildirim, A. Aharony, O. Entin-
Wohlman, and I. Ya. Korenblit, Unusual Symmetries in
the Kugel-Khomskii Hamiltonian , Phys. Rev. Lett. 91,
087206 (2003).
[79] M. Daghofer, K. Wohlfeld, A. M. Ole s, E. Arrigoni, and
P. Horsch, Absence of Hole Connement in Transition-
Metal Oxides with Orbital Degeneracy , Phys. Rev. Lett.
100, 066403 (2008).
[80] A. M. Ole s, Antiferromagnetism and Correlation of Elec-
trons in Transition Metals , Phys. Rev. B 28, 327 (1983).
[81] M. Cuoco, F. Forte, and C. Noce, Interplay of Coulomb
Interactions and c-Axis Octahedra Distortions in Single-
Layer Ruthenates , Phys. Rev. B 74, 195124 (2006).
[82] P. Marra, K. Wohlfeld, and J. van den Brink, Unraveling
Orbital Correlations with Magnetic Resonant Inelastic X-
Ray Scattering , Phys. Rev. Lett. 109, 117401 (2012).
[83] M. A. Hossain, I. Zegkinoglou, Y.-D. Chuang, J. Geck,
B. Bohnenbuck, A. G. Cruz Gonzalez, H.-H. Wu,
C. Sch uler-Langeheine, D. G. Hawthorn, J. D. Den-
linger, R. Mathieu, Y. Tokura, S. Satow, H. Takagi,
Y. Yoshida, Z. Hussain, B. Keimer, G. A. Sawatzky, and
A. Damascelli, Electronic Superlattice Revealed by Reso-
nant Scattering from Random Impurities in Sr3Ru2O7,
Sci. Rep. 3, 2299 (2013).
[84] L. Limot, P. Mendels, G. Collin, C. Mondelli, B. Oulad-
diaf, H. Mutka, N. Blanchard, and M. Mekata, Suscep-
tibility and Dilution Eects of the kagom e Bilayer Ge-
ometrically Frustrated Network: A Ga NMR study of
SrCr 9pGa129pO19, Phys. Rev. B 65, 144447 (2002).
[85] J. Bobro, N. La
orencie, L. K. Alexander, A. V. Maha-
jan, B. Koteswararao, and P. Mendels, Impurity-Induced
Magnetic Order in Low-Dimensional Spin-Gapped Mate-
rials, Phys. Rev. Lett. 103, 047201 (2009).
[86] A. Sen, K. Damle, and R. Moessner, Fractional Spin Tex-
tures in the Frustrated Magnet SrCr 9pGa129pO19, Phys.
Rev. Lett. 106, 127203 (2011).
[87] P. Bonf a, P. Carretta, S. Sanna, G. Lamura, G. Prando,
A. Martinelli, A. Palenzona, M. Tropeano, M. Putti, and
R. De Renzi, Magnetic Properties of Spin-diluted Iron
Pnictides from SR and NMR in LaFe 1xRuxAsO, Phys.
Rev. B 85, 054518 (2012).
[88] A. K. Paul, M. Reehuis, V. Ksenofontov, B. Yan,
A. Hoser, D. M. T obbens, P. M. Abdala, P. Adler,
M. Jansen, and C. Felser, Lattice Instability and Com-
peting Spin Structures in the Double Perovskite Insulator
Sr2FeOsO 6, Phys. Rev. Lett. 111, 167205 (2013). |
2102.01400v2.Coupled_spin_orbital_fluctuations_in_a_three_orbital_model_for__4d__and__5d__oxides_with_electron_fillings__n_3_4_5_____Application_to___rm_NaOsO_3_____rm_Ca_2RuO_4___and___rm_Sr_2IrO_4_.pdf | arXiv:2102.01400v2 [cond-mat.str-el] 20 Mar 2021Coupled spin-orbital fluctuations in a three orbital model f or4d
and5doxides with electron fillings n= 3,4,5— Application to
NaOsO 3, Ca2RuO4, and Sr 2IrO4
Shubhajyoti Mohapatra and Avinash Singh∗
Department of Physics, Indian Institute of Technology, Kanpu r - 208016, India
(Dated: March 23, 2021)
A unified approach is presented for investigating coupled sp in-orbital fluctuations
within a realistic three-orbital model for strongly spin-o rbit coupled systems with
electron fillings n= 3,4,5 in the t2gsector of dyz,dxz,dxyorbitals. A generalized
fluctuation propagator is constructed which is consistent w ith the generalized self-
consistent Hartree-Fock approximation where all Coulomb i nteraction contributions
involving orbital diagonal and off-diagonal spin and charge c ondensates are included.
Besides the low-energy magnon, intermediate-energy orbit on and spin-orbiton, and
high-energy spin-orbit exciton modes, the generalized spe ctral function also shows
other high-energy excitations such as the Hund’s coupling i nduced gapped magnon
modes. We relate the characteristic features of the coupled spin-orbital excitations
to the complex magnetic behavior resulting from the interpl ay between electronic
bands, spin-orbit coupling, Coulomb interactions, and str uctural distortion effects,
as realized in the compounds NaOsO 3, Ca2RuO4, and Sr 2IrO4.2
I. INTRODUCTION
The 4dand 5dtransition metal (TM) oxides exhibit an unprecedented coupling bet ween
spin, charge, orbital, and structural degrees of freedom. The c omplex interplay between
the different physical elements such as strong spin-orbit coupling ( SOC), Coulomb interac-
tions, and structural distortions results in novel magnetic state s and unconventional collec-
tive excitations.1–6In particular, the cubic structured NaOsO 3and perovskite structured
Ca2RuO4and Sr 2IrO4compounds, corresponding to dnelectronic configuration of the TM
ion with electron fillings n=3,4,5 in the t 2gsector, respectively, are at the emerging research
frontier as they provide versatile platform for the exploration of S OC-driven phenomena
involving collective electronic and magnetic behavior including coupled s pin-orbital excita-
tions.
Thedifferent physical elements giverisetoarichvarietyofnontrivia l microscopic features
which contribute to the complex interplay. These include spin-orbita l-entangled states,
bandnarrowing, spin-orbit gap, andexplicit spin-rotation-symmet ry breaking (dueto SOC),
electronic band narrowing due to reduced effective hopping (octah edral tilting and rotation),
crystal field induced tetragonal splitting (octahedral compress ion), orbital mixing (SOC
and octahedral tilting, rotation) which self consistently generate s induced SOC terms and
orbitalmomentinteractionfromtheCoulombinteractionterms, sig nificantlyweakerelectron
correlation term Ucompared to 3 dorbitals and therefore critical contribution of Hund’s
coupling to local magnetic moment. These microscopic features con tribute to the complex
interplay in different ways for electron fillings n=3,4,5, resulting in significantly different
macroscopic properties of the three compounds, which are briefly reviewed below along
with experimental observations about the collective and coupled sp in-orbital excitations as
obtained from recent resonant inelastic X-ray scattering (RIXS) studies.
The nominally orbitally quenched d3compound NaOsO 3undergoes a metal-insulator
transition (MIT) ( TMI=TN= 410 K) that is closely related to the onset of long-range
antiferromagnetic (AFM) order.7–10Various mechanisms, such as Slater-like, magnetic Lif-
shitz transition, and AFM band insulator have been proposed to exp lain this unusual and
intriguing nature of the MIT.8,11–14Interplay of electronic correlations, Hund’s coupling,
and octahedral tilting and rotation induced band narrowing near th e Fermi level in this
weakly correlated compound results in the weakly insulating state wit h G-type AFM or-3
der, with magnetic anisotropy and large magnon gap resulting from in terplay of SOC, band
structure, and the tetragonal splitting.14,15The OsL3resonant edge RIXS measurements
at room temperature show four inelastic peak features below 1.5 eV , which have been in-
terpreted to correspond to the strongly gapped ( ∼58 meV) dispersive magnon excitations
with bandwidth ∼100 meV, excitations (centered at ∼1 eV) within the t2gmanifold, and
excitations from t2gtoegstates and ligand-to-metal charge transfer for the remaining tw o
higher-energypeaks.13,16,17Theintensity andpositionsofthethreehigh-energypeaksappear
to be essentially temperature independent.
The nominally spin S=1d4compound Ca 2RuO4undergoes a MIT at TMI=357 K and
magnetic transition at TN=110 K ( ≪TMI) via a structural phase transition involving a
compressive tetragonal distortion, tilt, and rotation of the RuO 6octahedra.18–21The low-
temperature AFM insulating phase is thus characterized by highly dis torted octahedra
with nominally filled xyorbital and half-filled yz,xzorbitals.22–24This transition has also
been identified in pressure,25–27chemical substitution,28–30strain,31and electrical current
studies,32,33and highlights the complex interplay between SOC, Coulomb interactio ns, and
structural distortions.
Inelastic neutron scattering (INS)34–36and Raman37studies on Ca 2RuO4have revealed
unconventional low-energy ( ∼50 and 80 meV) excitations interpreted as gapped trans-
verse magnon modes and possibly soft longitudinal (“Higgs-like”) or two-magnon excitation
modes. From both Ru L3-edge and oxygen K-edge RIXS studies, multiple nontrivial exci-
tations within the t2gmanifold were observed recently below 1 eV.38–40Two low-energy ( ∼
80 and 350 meV) and two high-energy ( ∼750 meV and 1 eV) excitations were identified
within the limited energy resolution of RIXS. From the incident angle an d polarisation de-
pendence of the RIXS spectra, the orbital character of the 80 m eV peak was inferred to be
mixture of xyandxz/yzstates, whereas the 0.4 eV peak was linked to unoccupied xz/yz
states. Guided by phenomenological spin models, the low-energy ex citations (consisting of
multiple branches) were interpreted as composite spin-orbital exc itations (also termed as
“spin orbitons”).
Finally, SOC induced novel Mott insulating state is realized in the d5compound
Sr2IrO4,41,42where band narrowing of the spin-orbital-entangled electronic sta tes near the
Fermi level plays a critical role in the insulating behavior. The AFM insu lating ground
state is characterized by the correlation induced insulating gap with in the nominally J=1/24
bandsemerging fromtheKramersdoublet, which areseparatedfr omthebands ofthe J=3/2
quartet by energy 3 λ/2, whereλis the SOC strength. The RIXS spectra show low-energy
dispersive magnon excitations (up to 200 meV), further resolved in to two gapped magnon
modes with energy gaps ∼40 meV and 3 meV at the Γ point corresponding to out-of-
plane and in-plane fluctuation modes, respectively.43–46Weak electron correlation effect and
mixing between the J=1/2 and 3/2 sectors were identified as contributing significantly to
the strong zone-boundary magnon dispersion as measured in RIXS studies.47In addition,
high-energy dispersive spin-orbit exciton modes have also been rev ealed in RIXS studies in
the energy range 0.4-0.8 eV.48This distinctive mode is also referred to as the spin-orbiton
mode,49,50and has been attributed to the correlated motion of electron-hole pair excitations
across the renormalized spin-orbit gap between the J=1/2 and 3/2 bands.51
Most of the theoretical studies involving magnetic anisotropy effec ts and excitations in
abovesystems havemainlyfocusedonphenomenological spinmodels withdifferent exchange
interactions obtained as fitting parameters to the experimental s pectra. However, the in-
terpretation of experimental data remains incomplete since the ch aracter of the effective
spins, the microscopic origin of their interactions, and the microsco pic nature of the mag-
netic excitations are still debated.2–6Realistic information about the spin-orbital character
of both low and high-energy collective excitations, as inferred from the study of coupled
spin-orbital excitations, is clearly important since the spin and orbit al degrees of freedom
are explicitly coupled, and both are controlled by the different physic al elements such as
SOC, Coulomb interaction terms, tetragonal compression induced crystal-field splitting be-
tweenxyandyz,xzorbitals, octahedral tilting and rotation induced orbital mixing hopp ing
terms, and band physics.
Due to the intimately intertwined roles of the different physical eleme nts, a unified ap-
proach is therefore required for the realistic modeling of these sys tems in which all physical
elements are treated on an equal footing. The generalized self-co nsistent approximation ap-
pliedrecentlytothe n= 4compoundCa 2RuO4providessuch aunifiedapproach.52Involving
theself-consistent determinationofmagneticorderwithinathree -orbitalinteractingelectron
model including all orbital-diagonal and off-diagonal spin and charge condensates generated
by the different Coulomb interaction terms, this approach explicitly in corporates the com-
plex interplay and accounts for the observed behavior including the tetragonal distortion
induced magnetic reorientation transition, orbital moment interac tion induced orbital gap,5
SOC and octahedral tilting induced easy-axis anisotropy, and Coulo mb interaction induced
anisotropic SOC renormalization. Extension to the n= 5 compound Sr 2IrO4,53provides
confirmation of the Hund’s coupling induced easy-plane magnetic anis otropy, which is re-
ponsible for the ∼40 meV magnon gap measured for the out-of-plane fluctuation mod e.46
Towards a generalized non-perturbative formalism unifying the mag netic order and
anisotropy effects on one hand and collective excitations on the oth er, the natural exten-
sion of the above generalized condensate approach is therefore t o consider the generalized
fluctuation propagator in terms of the generalized spin ( ψ†
µ[σα]ψν) and charge ( ψ†
µ[1]ψν) op-
erators in the pure spin-orbital basis of the t 2gorbitalsµ,ν=yz,xz,xy and spin components
α=x,y,z. The generalized operators include the normal ( µ=ν) spin and charge opera-
tors as well as the orbital off-diagonal ( µ/negationslash=ν) cases which are related to the generalized
spin-orbit coupling terms ( LαSβ, whereα,β=x,y,z) and the orbital angular momentum
operatorsLα. Constructing the generalized fluctuation propagator as above w ill ensure that
this scheme is fully consistent with the generalized self-consistent a pproach involving the
generalized condensates.
The different components of the generalized fluctuation propagat or will therefore natu-
rallyincludespin-orbitonsandorbitons, corresponding tothespin- orbital(LαSβ)andorbital
(Lα) moment fluctuations, besides the normal spin andchargefluctua tions. The normal spin
fluctuations will include in-phase and out-of phase fluctuations with respect to different or-
bitals, the latter being strongly gapped due to Hund’s coupling. The s pin-orbitons will
include the spin-orbit excitons measured in RIXS studies of Sr 2IrO4.
The structure of this paper is as below. The three-orbital model w ithin the t 2gsector
(including SOC, hopping, Coulomb interaction, and structural disto rtion terms), and the
generalized self-consistent formalism including orbital diagonal and off-diagonal condensates
are reviewed in Sec. II and III. After introducing the generalized fl uctuation propagator in
Sec. IV, results of the calculated fluctuation spectral functions are presented for the cases
n= 3,4,5 (corresponding to the three compounds NaOsO 3, Ca2RuO4, Sr2IrO4) in Sections
V, VI, VII. Finally, conclusions are presented in Sec. VIII. The bas is-resolved contributions
to the total spectral function showing the detailed spin-orbital c haracter of the collective
excitations are presented in the Appendix.6
II. THREE ORBITAL MODEL WITH SOC AND COULOMB INTERACTIONS
In the three-orbital ( µ=yz,xz,xy ), two-spin ( σ=↑,↓) basis defined with respect to a
common spin-orbital coordinate axes (Fig. 1), we consider the Ham iltonianH=Hband+
Hcf+Hint+HSOCwithin the t2gmanifold. For the band and crystal field terms together,
we consider:
Hband+cf=/summationdisplay
kσsψ†
kσs
ǫyz
k′0 0
0ǫxz
k′0
0 0ǫxy
k′+ǫxy
δss′+
ǫyz
kǫyz|xz
kǫyz|xy
k
−ǫyz|xz
kǫxz
kǫxz|xy
k
−ǫyz|xy
k−ǫxz|xy
kǫxy
k
δ¯ss′
ψkσs′
(1)
in the composite three-orbital, two-sublattice ( s,s′= A,B) basis. Here the energy offset
ǫxy(relative to the degenerate yz/xzorbitals) represents the tetragonal distortion induced
crystal field effect. The band dispersion terms in the two groups co rrespond to hopping
terms connecting the same and opposite sublattice(s), and are giv en by:
ǫxy
k=−2t1(coskx+cosky)
ǫxy
k′=−4t2coskxcosky−2t3(cos2kx+cos2ky)
ǫyz
k=−2t5coskx−2t4cosky
ǫxz
k=−2t4coskx−2t5cosky
ǫyz|xz
k=−2tm1(coskx+cosky)
ǫxz|xy
k=−2tm2(2coskx+cosky)
ǫyz|xy
k=−2tm3(coskx+2cosky). (2)
Heret1,t2,t3are respectively the first, second, and third neighbor hopping ter ms for
thexyorbital. For the yz(xz) orbital,t4andt5are the nearest-neighbor (NN) hopping
terms iny(x) andx(y) directions, respectively, corresponding to πandδorbital overlaps.
Octahedral rotation and tilting induced orbital mixings are represe nted by the NN hopping
termstm1(betweenyzandxz) andtm2,tm3(betweenxyandxz,yz). In then= 4 case
corresponding to the Ca 2RuO4compound, we have taken hopping parameter values: ( t1,t2,
t3,t4,t5)=(−1.0,0.5,0,−1.0,0.2),orbital mixing hopping terms: tm1=0.2andtm2=tm3=0.15
(≈0.2/√
2), andǫxy=−0.8, all in units of the realistic hopping energy scale |t1|=150
meV.54–56The choice tm2=tm3corresponds to the octahedral tilting axis oriented along7
(a)(b)
FIG. 1: (a) The common spin-orbital coordinate axes ( x−y) along the Ru-O-Ru directions, shown
along with the crystal axes a,b. (b) Octahedral tilting about the crystal aaxis is resolved along
thex,yaxes, resulting in orbital mixing hopping terms between the xyandyz,xzorbitals.
the±(−ˆx+ ˆy) direction, which is equivalent to the crystal ∓adirection (Fig. 1). The tm1
andtm2,m3values taken above approximately correspond to octahedral rot ation and tilting
angles of about 12◦(≈0.2 rad) as reported in experimental studies.26
For the on-site Coulomb interaction terms in the t2gbasis (µ,ν=yz,xz,xy ), we consider:
Hint=U/summationdisplay
i,µniµ↑niµ↓+U′/summationdisplay
i,µ<ν,σniµσniνσ+(U′−JH)/summationdisplay
i,µ<ν,σniµσniνσ
+JH/summationdisplay
i,µ/negationslash=νa†
iµ↑a†
iν↓aiµ↓aiν↑+JP/summationdisplay
i,µ/negationslash=νa†
iµ↑a†
iµ↓aiν↓aiν↑
=U/summationdisplay
i,µniµ↑niµ↓+U′′/summationdisplay
i,µ<νniµniν−2JH/summationdisplay
i,µ<νSiµ.Siν+JP/summationdisplay
i,µ/negationslash=νa†
iµ↑a†
iµ↓aiν↓aiν↑(3)
including the intra-orbital ( U) and inter-orbital ( U′) density interaction terms, the Hund’s
coupling term ( JH), and the pair hopping interaction term ( JP), withU′′≡U′−JH/2 =
U−5JH/2 from the spherical symmetry condition U′=U−2JH. Herea†
iµσandaiµσare
the electron creation and annihilation operators for site i, orbitalµ, spinσ=↑,↓. The
density operator niµσ=a†
iµσaiµσ, total density operator niµ=niµ↑+niµ↓=ψ†
iµψiµ, and spin
density operator Siµ=ψ†
iµσψiµin terms of the electron field operator ψ†
iµ= (a†
iµ↑a†
iµ↓). All
interaction terms above are SU(2) invariant and thus possess spin rotation symmetry.
Finally, for the bare spin-orbit coupling term (for site i), we consider the spin-space8
representation:
HSOC(i) =−λL.S=−λ(LzSz+LxSx+LySy)
=
/parenleftig
ψ†
yz↑ψ†
yz↓/parenrightig/parenleftig
iσzλ/2/parenrightig
ψxz↑
ψxz↓
+/parenleftig
ψ†
xz↑ψ†
xz↓/parenrightig/parenleftig
iσxλ/2/parenrightig
ψxy↑
ψxy↓
+/parenleftig
ψ†
xy↑ψ†
xy↓/parenrightig/parenleftig
iσyλ/2/parenrightig
ψyz↑
ψyz↓
+H.c. (4)
which explicitly breaks SU(2) spin rotation symmetry and therefore generates anisotropic
magnetic interactions from its interplay with other Hamiltonian terms . Here we have used
the matrix representation:
Lz=
0−i0
i0 0
0 0 0
, Lx=
0 0 0
0 0−i
0i0
, Ly=
0 0i
0 0 0
−i0 0
, (5)
for the orbital angular momentum operators in the three-orbital (yz,xz,xy ) basis.
As the orbital “hopping” terms in Eq. (4) have the same form as spin -dependent hopping
termsiσ.t′
ij, carrying out the strong-coupling expansion57for the−λLzSzterm to second
order inλyields the anisotropic diagonal (AD) intra-site interactions:
[H(2)
eff](z)
AD(i) =4(λ/2)2
U/bracketleftbig
Sz
yzSz
xz−(Sx
yzSx
xz+Sy
yzSy
xz)/bracketrightbig
(6)
betweenyz,xzmoments if these orbitals arenominally half-filled, as in thecase ofCa 2RuO4.
This term explicitly yields preferential x−yplane ordering (easy-plane anisotropy) for
parallelyz,xzmoments, as enforced by the relatively stronger Hund’s coupling.
Similarly, from the strong coupling expansion for the other two SOC t erms, we obtain
additional anisotropic interaction terms which are shown below to yie ldC4symmetric easy-
axis anisotropy within the easy plane. From the −λLxSxand−λLySyterms, we obtain:
[H(2)
eff](x,y)
AD(i) =4(λ/2)2
U/bracketleftbig
Sx
xzSx
xy−(Sy
xzSy
xy+Sz
xzSz
xy)/bracketrightbig
+4(λ/2)2
U/bracketleftbig
Sy
xySy
yz−(Sx
xySx
yz+Sz
xySz
yz)/bracketrightbig
(7)
Neglecting the terms involving the Szcomponents which are suppressed by the easy-plane
anisotropy discussed above, we obtain:
[H(2)
eff](x,y)
AD(i) =−4(λ/2)2
U/bracketleftbig
Sx
xy(Sx
yz−Sx
xz)+Sy
xy(Sy
xz−Sy
yz)/bracketrightbig
=−4(λ/2)2
UfxyS2[sin2φsinφc] (8)9
where the spin components are expressed as: Sx
xy=fxyScosφ,Sx
yz=Scos(φ−φc),Sx
xz=
Scos(φ+φc) (and similarly for the ycomponents) in terms of the overall orientation angle φ
of the magnetic order and the relative canting angle 2 φcbetween the yz,xzmoments. Here
the factorfxy<1 represents the reduced moment for the xyorbital.
The above expression shows the composite orientation and canting angle dependence of
the anisotropic interaction energy having the C4symmetry. Minimum energy is obtained
at orientations φ=nπ/4 (wheren= 1,3,5,7) since the canting angle has the approximate
functional form φc≈φmax
csin2φin terms of the orientation φ. Thus, while the easy-plane
anisotropy involves only the yz,xzmoments, the xymoment plays a crucial role in the
easy-axis anisotropy, which is directly relevant for NaOsO 3(xyorbital is also nominally
half-filled), but also for Ca 2RuO4with the factor fxyas incorporated above.
For later reference, we note here that condensates of the orbit al off-diagonal (OOD) one-
body operators as in Eq. (4) directly yield physical quantities such a s orbital magnetic
moments and spin-orbital correlations:
/angbracketleftLα/angbracketright=−i/bracketleftbig
/angbracketleftψ†
µψν/angbracketright−/angbracketleftψ†
µψν/angbracketright∗/bracketrightbig
= 2 Im/angbracketleftψ†
µψν/angbracketright
/angbracketleftLαSα/angbracketright=−i/bracketleftbig
/angbracketleftψ†
µσαψν/angbracketright−/angbracketleftψ†
µσαψν/angbracketright∗/bracketrightbig
/2 = Im/angbracketleftψ†
µσαψν/angbracketright
λint
α= (U′′−JH/2)/angbracketleftLαSα/angbracketright= (U′′−JH/2)Im/angbracketleftψ†
µσαψν/angbracketright (9)
where the orbital pair ( µ,ν) corresponds to the component α=x,y,z, and the last equation
yields the interaction induced SOC renormalization, as discussed in th e next section.
III. SELF-CONSISTENT DETERMINATION OF MAGNETIC ORDER
We consider the various contributions from the Coulomb interaction terms (Eq. 3) in the
HF approximation, focussing first on terms with normal (orbital dia gonal) spin and charge
condensates. The resulting local spin and charge terms can be writ ten as:
[HHF
int]normal=/summationdisplay
iµψ†
iµ[−σ.∆iµ+Eiµ1]ψiµ (10)
where the spin and charge fields are self-consistently determined f rom:
2∆α
iµ=U/angbracketleftσα
iµ/angbracketright+JH/summationdisplay
ν<µ/angbracketleftσα
iν/angbracketright(α=x,y,z)
Eiµ=U/angbracketleftniµ/angbracketright
2+U′′/summationdisplay
ν<µ/angbracketleftniν/angbracketright (11)10
in terms of the local charge density /angbracketleftniµ/angbracketrightand the spin density components /angbracketleftσα
iµ/angbracketright.
There are additional contributions resulting from orbital off-diago nal (OOD) spin and
charge condensates which are finite due to orbital mixing induced by SOC and structural
distortions (octahedral tilting and rotation). The contributions c orresponding to different
Coulomb interaction terms are summarized in Appendix A, and can be g rouped in analogy
with Eq. (10) as:
[HHF
int]OOD=/summationdisplay
i,µ<νψ†
iµ[−σ.∆iµν+Eiµν1]ψiν+H.c. (12)
where the orbital off-diagonal spin and charge fields are self-cons istently determined from:
∆iµν=/parenleftbiggU′′
2+JH
4/parenrightbigg
/angbracketleftσiνµ/angbracketright+/parenleftbiggJP
2/parenrightbigg
/angbracketleftσiµν/angbracketright
Eiµν=/parenleftbigg
−U′′
2+3JH
4/parenrightbigg
/angbracketleftniνµ/angbracketright+/parenleftbiggJP
2/parenrightbigg
/angbracketleftniµν/angbracketright (13)
in terms of the corresponding condensates /angbracketleftσiµν/angbracketright ≡ /angbracketleftψ†
iµσψiν/angbracketrightand/angbracketleftniµν/angbracketright ≡ /angbracketleftψ†
iµ1ψiν/angbracketright.
The spin andcharge condensates inEqs. 11and 13 areevaluated us ing the eigenfunctions
(φk) and eigenvalues ( Ek) of the full Hamiltonian in the given basis including the interaction
contributions [ HHF
int] (Eqs. 10 and 12) using:
/angbracketleftσα
iµν/angbracketright ≡ /angbracketleftψ†
iµσαψiν/angbracketright=Ek<EF/summationdisplay
k(φ∗
kµs↑φ∗
kµs↓)[σα]
φkνs↑
φkνs↓
(14)
for siteion thes=A/Bsublattice, and similarly for the charge condensates /angbracketleftniµν/angbracketright ≡
/angbracketleftψ†
iµ1ψiν/angbracketrightwith the Pauli matrices [ σα] replaced by the unit matrix [ 1]. The normal spin and
charge condensates correspond to ν=µ. For each orbital pair ( µ,ν) = (yz,xz), (xz,xy),
(xy,yz), there are three components ( α=x,y,z) for the spin condensates /angbracketleftψ†
µσαψν/angbracketrightand
one charge condensate /angbracketleftψ†
µ1ψν/angbracketright. This is analogous to the three-plus-one normal spin and
charge condensates for each of the three orbitals µ=yz,xz,xy .
The above additional terms involving orbital off-diagonal condensa tes contribute to or-
bital physics. Thus, the charge terms lead to coupling of orbital an gular momentum opera-
tors to weak orbital fields, the spin terms result in interaction-indu ced SOC renormalization
as given in Eq. (9), and the self consistently determined renormalize d SOC values are
obtained as:
λα=λ+λint
α (15)11
forthethreecomponents α=x,y,z. Resultsoftheselfconsistentdeterminationofmagnetic
order including all orbital diagonal and off-diagonal spin and charge condensates have been
presented for the 4d4compound Ca 2RuO4recently,52illustrating the rich interplay between
different physical elements.
IV. GENERALIZED FLUCTUATION PROPAGATOR
Sinceallgeneralizedspin /angbracketleftψ†
µσψν/angbracketrightandcharge /angbracketleftψ†
µψν/angbracketrightcondensateswereincludedintheself
consistent determination of magnetic order, the fluctuation prop agator must also be defined
in terms of the generalized operators. We therefore consider the time-ordered generalized
fluctuation propagator:
[χ(q,ω)] =/integraldisplay
dt/summationdisplay
ieiω(t−t′)e−iq.(ri−rj)×/angbracketleftΨ0|T[σα
µν(i,t)σα′
µ′ν′(j,t′)]|Ψ0/angbracketright (16)
in the self-consistent AFM ground state |Ψ0/angbracketright, where the generalized spin-charge operators
at lattice sites i,jare defined as σα
µν=ψ†
µσαψν, which include both the orbital diagonal
(µ=ν) and off-diagonal ( µ/negationslash=ν) cases, as well as the spin ( α=x,y,z) and charge ( α=c)
operators, with σαdefined as Pauli matrices for α=x,y,zand unit matrix for α=c.
In the random phase approximation (RPA), the generalized fluctua tion propagator is
obtained as:
[χ(q,ω)]RPA=2[χ0(q,ω)]
1−[U][χ0(q,ω)](17)
in terms of the bare particle-hole propagator [ χ0(q,ω)] which is evaluated by integrating out
the electronic degrees of freedom:
[χ0(q,ω)]µ′ν′α′s′
µναs=1
2/summationdisplay
k/bracketleftigg
/angbracketleftk|σα
µν|k−q/angbracketrights/angbracketleftk|σα′
µ′ν′|k−q/angbracketright∗
s′
E⊕
k−q−E⊖
k+ω−iη+/angbracketleftk|σα
µν|k−q/angbracketrights/angbracketleftk|σα′
µ′ν′|k−q/angbracketright∗
s′
E⊕
k−E⊖
k−q−ω−iη/bracketrightigg
(18)
The matrix elements in the above expression are evaluated using the eigenvectors of the HF
Hamiltonian in the self-consistent AFM state:
/angbracketleftk|σα
µν|k−q/angbracketrights= (φ∗
kµ↑sφ∗
kµ↓s)[σα]
φk−qν↑s
φk−qν↓s
(19)12
and the superscripts ⊕(⊖) refer to particle (hole) states above (below) the Fermi energy.
The subscripts s,s′indicate thetwo (A/B) sublattices. Inthe compositespin-charge- orbital-
sublattice ( µναs) basis, the [ χ0(q,ω)] matrix is of order 72 ×72, and the form of the [ U]
matrix in the RPA expression (Eq. 17) is given in Appendix B.
The spectral function of the excitations will be determined from:
Aq(ω) =1
πIm Tr[χ(q,ω)]RPA (20)
using the RPA expression for [ χ(q,ω)]. When the collective excitation energies lie within
the AFM band gap, it is convenient to consider the symmetric form of the denominator in
the RPA expression (Eq. 17):
[U][χ0(q,ω)][U]−[U] (21)
and in terms of the real eigenvalues λq(ω) of this Hermitian matrix, the magnon energies
ωqfor momentum qare determined by solving for the zeroes:
λq(ω=ωq) = 0 (22)
corresponding to the poles in the propagator.
Results of the calculated spectral function will be discussed in the s ubsequent sections
for different electron filling cases ( n= 3,4,5) with applications to corresponding 4 dand
5dtransition metal compounds. Broadly, our investigation of the gen eralized fluctuation
propagator will provide information about the dominantly spin, orbit al, and spin-orbital
excitations, as the generalized spin and charge operators ψ†
µσαψνinclude spin ( µ=ν,
α=x,y,z), orbital (µ/negationslash=ν,α=c), and spin-orbital ( µ/negationslash=ν,α=x,y,z) cases. Also
included will be the high-energy spin-orbit exciton modes involving par ticle-hole excitations
across the renormalized spin-orbit gap between spin-orbital enta ngled states of different J
sectors, as in the n= 5 case relevant for the Sr 2IrO4compound.
V.n= 3— APPLICATION TO NaOsO 3
The strongly spin-orbit coupled orthorhomic structured 5 d3osmium compound NaOsO 3,
with nominally three electrons in the Os t2gsector, exhibits several novel electronic and
magnetic properties. These include a G-type antiferromagnetic (A FM) structure with spins13
oriented along the caxis, a significantly reduced magnetic moment ∼1µBas measured from
neutron scattering, a continuous metal-insulator transition (MIT ) that coincides with the
AFM transition ( TN=TMIT= 410 K) as seen in neutron and X-ray scattering, and a large
magnon gap of 58 meV as seen in resonant inelastic X-ray scattering (RIXS) measurements
indicating strong magnetic anisotropy.8–10,16
Two different mechanisms contributing to SOC-induced easy-plane a nisotropy and large
magnon gap for out-of-plane fluctuation modes were identified for the weakly correlated
5d3compound NaOsO 3in terms of a simplified picture involving only the normal spin and
charge densities.14,15Both essential ingredients — (i) small moment disparity between yz,xz
andxyorbitals and (ii) spin-charge coupling effect in presence of tetragon al splitting — are
intrinsically present in the considered three-orbital model on the s quare lattice. A realistic
representation of magnetic anisotropy in NaOsO 3is therefore provided by the considered
model, while maintaining uniformity of lattice structure across the n= 3,4,5 cases consid-
ered in order to keep the focus on coupled spin-orbital fluctuation s.
The first mechanism involves the SOC-induced anisotropic interactio n terms as in Eq.
(6) resulting from the three SOC terms −λLαSαforα=x,y,z. Due to the small moment
disparitymyz,xz>mxyresulting from the broader xyband, the interaction term in Eq. (6)
dominates over the other two terms, leading to the easy-plane anis otropy for parallel yz,xz
moments enforced by the Hund’s coupling. With increasing U, this effect weakens as the
moments saturate myz,xz,xy≈1 in the large Ulimit. In the second mechanism, the SOC
induced decreasing xyorbital density nxywith spin rotation from zdirection to x−yplane
couples to the tetragonal distortion term, and for positive ǫxythe energy is minimized for
spin orientation in the x−yplane.
We will consider the parameter set values U= 4,JH=U/5,U′′=U−5JH/2, bare
SOC value λ=1.0, andǫxy= 0.5 unless otherwise indicated, with the hopping energy scale
|t1|=300 meV. Thus, U= 1.2 eV,λ=0.3 eV,ǫxy= 0.15 eV, which are realistic values for the
NaOsO 3compound. Initially, we will also set tm1,m2,m3= 0 for simplicity, and focus on the
easy-plane anisotropy and large magnon gap for out-of-plane fluc tuations.
Self consistent determination of magnetic order using the generaliz ed approach discussed
in Sec. III confirms the easy-plane anisotropy. Starting in nearly zdirection, the AFM order
direction self consistently approaches the x−yplane in a few hundred iterations. Initially,
we will discuss magnetic excitations in the self consistent state with A FM order along the14
(π/2,π/2) (π,0) ( π,π) (0,0) 0 50 100 150 200 250 300ω (meV)
0.01 0.1 1 10 100
(a)
0 50 100 150 200 250 300
(π/2,π/2) (π,0) ( π,π) (π/2,π/2)(b)ωq (meV)
FIG. 2: (a) Low energy part of the calculated spectral functi on from the generalized fluctuation
propagator shows the magnon excitations in the self-consis tent state with planar AFM order, and
(b) magnon dispersion showing the gapless and gapped modes c orresponding to in-plane and out-
of-plane fluctuations.
ˆxor ˆydirections. Although these orientations correspond to metastab le states as discussed
later, they provide convenient test cases for explicitly confirming t he gapless in-plane and
gapped out-of-plane magnon modes in the generalized fluctuation p ropagator calculation.
The low-energy part of the calculated spectral function using Eq. (20) is shown in the
Fig. 2(a) as an intensity plot for qalong symmetry directions of the Brillouin zone. The
gapless and gapped modes corresponding to in-plane and out-of-p lane fluctuations reflect
the easy-plane magnetic anisotropy. The calculated gap energy 60 meV is close to the
measured spin wave gap of 58 meV in NaOsO 3. Also shown for comparison in Fig. 2(b)
is the magnon dispersion calculated from the poles of the RPA propag ator as described in
Sec. IV. Focussing on the magnon gap in Fig. 2(b), which provides a m easure of the SOC
induced easy-plane anisotropy, effects of various physical quant ities are shown in Fig. 3.
ThegaplessGoldstonemodecorrespondingtoin-planerotationofA FMorderingdirection
in thex−yplane involves only small changes in spin densities /angbracketleftψ†
µσαψµ/angbracketrightforα=x,yand
µ=yz,xz, and also in generalized spin densities /angbracketleftψ†
µσαψν/angbracketrightforα=x,yandµ=yz,xzwith
fixedν=xy. For example, the magnetization values mx
yz= 0.82 andmx
xz= 0.84 change
tomy
yz= 0.84 andmy
xz= 0.82 when the ordering is rotated from xtoydirection. Thus,
the Goldstone mode is nearly pure spin mode and the small orbital cha racter reflects the
effectively suppressed spin-orbital entangement in the n= 3 AFM state. In contrast, the15
0 20 40 60 80 100
0 0.06 0.12 0.18 0.24 0.30 0.36(a)magnon gap (meV)
SOC (eV) 0 20 40 60 80 100
0.9 1.2 1.5 1.8 2.1 2.4(b)magnon gap (meV)
U (eV) 0 20 40 60 80 100
-0.15-0.1-0.05 0 0.05 0.1 0.15 0.2 0.25(c)magnon gap (meV)
εxy (eV)
FIG. 3: Variation of the calculated magnon gap showing effects of (a) SOC, (b) Hubbard U, and
(c) tetragonal distortion ǫxy, on the easy-plane magnetic anisotropy.
n= 5 case corresponding to Sr 2IrO4shows strongly coupled spin-orbital character of the
Goldstone mode (Appendix C) due to the extreme spin-orbital enta nglement.
We now consider the easy-axis anisotropy effects in our self consist ent determination of
magnetic order. With respect to the AFM order orientation (azimut hal angleφ) within the
easy (x−y) plane, we find an easy-axis anisotropy along the diagonal orientat ionsφ=nπ/4
(n= 1,3,5,7) even for no octahedral tilting. This anisotropy is due to the orien tation
and canting angle dependent anisotropic interaction (Eq. 8) as disc ussed in Sec. II. The
anisotropic interaction energy vanishes for φalong thex,yaxes (hence the gapless in-plane
mode in Fig. 2), and is significant near the diagonal orientations, res ulting in easy-axis
anisotropy and small relative canting between yz,xzmoments which is explicitly confirmed
in our self-consistent calculation.
The resulting C4symmetry of the easy-axis ±(ˆx±ˆy) is reduced to C2symmetry ±(ˆx−ˆy)
in the presence of octahedral tilting. The important anisotropy eff ects of the octahedral
tilting induced inter-site DM interactions are discussed below. We find that the DM axis
lies along the crystal baxis, leading to easy axis direction along the crystal aaxis. Both
these directions are interchanged in comparison to the Ca 2RuO4case, which follows from a
subtle difference in the present n= 3 case as explained below.
Following the analysis carried out for the Ca 2RuO4compound,52within the usual strong-
coupling expansion in terms of the normal ( t) and spin-dependent ( t′
x,t′
y) hopping terms
induced by the combination of SOC and orbital mixing hopping terms tm2,m3due to octa-
hedral tilting, the DM interaction terms generated in the effective s pin model are obtained16
TABLE I: Self consistently determined magnetization and de nsity values for the three orbitals ( µ)
on the two sublattices ( s), showing easy-axis anisotropy along the crystal aaxis due to octahedral
tilting induced DM interaction. Here tm2,m3= 0.15.
µ(s)mx
µmy
µmz
µnµ
yz(A) 0.598 −0.557 0.006 1.012
xz(A) 0.557 −0.598−0.006 1.012
xy(A) 0.541 −0.541 0.0 0.977µ(s)mx
µmy
µmz
µnµ
yz(B)−0.598 0.557 0.006 1.012
xz(B)−0.557 0.598 −0.006 1.012
xy(B)−0.541 0.541 0.0 0.977
TABLE II: Self consistently determined renormalized SOC va luesλα=λ+λint
αand the orbital
magnetic moments /angbracketleftLα/angbracketrightforα=x,y,zon the two sublattices. Bare SOC value λ=1.0.
s λ xλyλz/angbracketleftLx/angbracketright /angbracketleftLy/angbracketright /angbracketleftLz/angbracketright
A 1.179 1.179 1.364 0.032 −0.032 0.0
B 1.179 1.179 1.364 −0.032 0.032 0.0
as:
[H(2)
eff](x,y)
DM=8tt′
x
U/summationdisplay
/angbracketlefti,j/angbracketrightxˆx.(Si,xz×Sj,xz)+8tt′
y
U/summationdisplay
/angbracketlefti,j/angbracketrightyˆy.(Si,yz×Sj,yz)
≈8t|t′
x|
U/summationdisplay
/angbracketlefti,j/angbracketright(ˆx+ ˆy).(Si,yz×Sj,yz) (23)
where we have taken t′
x=−t′
y=−ive andSx
i,xz=Sx
i,yz(due to Hund’s coupling) as earlier,
but withSz
i,xz=−Sz
i,yzfor then= 3 case as obtained in our self consistent calculation which
is discussed below. The effective DM axis (ˆ x+ ˆy) is thus along the crystal baxis (Fig. 1)
for theyzorbital, resulting in easy-axis anisotropy along the crystal adirection, as well as
spin canting about the DM axis in the zdirection.
Results for various physical quantities are shown in Tables I and II. Starting with initial
orientation along the ˆ xor ˆydirections, the AFM order direction self consistently approaches
the easy-axis direction in a few hundred iterations, explicitly exhibitin g the strong easy-axis
anisotropywithintheeasy( x−y)planeduetotheoctahedraltiltinginducedDMinteraction,17
-3-2-1 0 1 2
(0,0) ( π,0) (π,π) (0,0) (0, π) (π,0)yz xz xy Ek - EF (eV)
0 50 100 150 200 250 300ωq (meV)
(π/2,π/2) (π,0) ( π,π) (π/2,π/2)
FIG. 4: (a) Calculated orbital resolved electronic band str ucture in the self-consistent state with
AFMorderalongthecrystal aaxisduetooctahedraltiltinginducedDMinteraction. Here tm2,m3=
0.15. Colors indicate dominant orbital weight: red ( yz), green ( xz), blue ( xy). (b) Magnon
dispersion for the magnetic order as given in Table I, showin g that both in-plane and out-of-plane
modes are appreciably gapped due to the easy-axis and easy-p lane anisotropies.
along with small spin canting in the zdirection about the DM axis. The small moment
disparitymyz,xz>mxyand the negligible orbital moments can also be seen here explicitly.
The renormalized SOC strength λzis enhanced relative to the other two components, which
further reduces the SOC induced frustration in this system with no minally one electron in
each of the three orbitals.
Withoctahedral tilting included, theorbital resolved electronic ban dstructure intheself-
consistent AFM state (Fig. 4(a)) shows the AFM band gap between valence and conduction
bands, SOC induced orbital mixing and band splittings, the fine splittin g due to octahedral
tilting, andtheasymmetricbandwidthfor xyorbitalbandscharacteristicofthe2ndneighbor
hopping term t2which connects the same magnetic sublattice. The calculated magno n
dispersion evaluated using Eq. 22 is shown in Fig. 4(b). As expected, both in-plane and
out-of-plane magnon modes are gapped due to the easy-axis and e asy-plane anisotropies
discussed above.
The high energy part of the spectral function is shown in the series of panels in Fig. 5
for different SOC strengths. The two groups of modes here corre spond to: (i) the Hund’s
coupling induced gapped magnon modes for out-of-phase spin fluct uations (the two disper-
sive modes starting at energies 0.7 and 0.8 eV from the left edge in pan el (a)), and (ii)18
(π/2,π/2) (π,0) ( π,π) (0,0) 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4ω (eV)
0.01 0.1 1 10 100
λ = 0(a)
(π/2,π/2) (π,0) ( π,π) (0,0) 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4ω (eV)
0.01 0.1 1 10 100
λ = 0.5(b)
(π/2,π/2) (π,0) ( π,π) (0,0) 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4ω (eV)
0.01 0.1 1 10 100
λ = 1.0(c)
(π/2,π/2) (π,0) ( π,π) (0,0) 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4ω (eV)
0.01 0.1 1 10 100
λ = εxy = 0 (d)
FIG. 5: Gapped magnon modes and dominantly magnetic exciton modes for the n= 3 case seen in
the high-energy part of the spectral function calculated in the self consistent AFM state including
octahedral tilting, for different SOC ( λ) values shown in the panels.
the spin-orbiton modes (starting at energy below 0.6 eV) which are in ter-orbital magnetic
excitons corresponding to the lowest-energy particle-hole excita tions across the AFM band
gap involving yz/xzorbitals (particle) and xyorbital (hole) states (Fig. 4(a)). Through
the usual resonant scattering mechanism, these modes are pulled down in energy below the
continuum by the U′′interaction term, and form well defined propagating modes.
The spin-orbiton mode involving xyorbital shifts to higher energy when ǫxydecreases to
zero (panel (d)) which lowers the dominantly xyvalence band (Fig. 4(a)) and thus increases
the particle-hole excitation energy. The splitting of the exciton mod es in panel (c) is due to
the SOC induced splitting of electronic bands as seen in Fig. 4(a), whic h is then reflected
in the particle-hole excitation energies. The combination of orbitals f or these exciton modes
indicates that LxandLycomponents of the orbital angular momentum are involved in these
coupled spin-orbital fluctuations. There is an additional spin-orbit on mode involving only
yz,xzorbitals (and Lzcomponent) which is formed at higher energy near 0.8 eV (flat band
near the left edge iin panel (a)). With increasing SOC, the high-ener gy modes involving19
yz,xzorbitals acquire significant spin-orbit exciton character.
VI.n= 5— APPLICATION TO Sr2IrO4
The perovskite structured 5 d5compound Sr 2IrO4exhibits an AFM insulating state due
to strong SOC induced splitting of the t2gstates, with four electrons in the nominally
filled and non-magnetic J=3/2 sector and one electron in the nominally half filled and
magnetically active J=1/2 sector. The SOC induced splitting of 3 λ/2 between states of the
two total angular momentum sectors, strong spin-orbital entan glement, and band narrowing
of states in the J=1/2 sector, all of these play a crucial role in the stabilization of the AFM
insulator state. Both low-energy magnon excitations and high-ene rgy spin-orbit excitons
acrosstherenormalizedspin-orbit gaphave beenintensively studie dusing RIXSexperiments
and variety of theoretical approaches.43,46,47,51,53
In this case, we have taken realistic parameter values U= 3,JH=U/7, bare SOC
valueλ= 1.35, andǫxy=−0.5 for simplicity, along with hopping terms: ( t1,t2,t3,t4,
t5,tm1)=(-1.0, 0.5, 0.25, -1.0, 0.0, 0.2), all in units of the realistic hopping en ergy scale
|t1|=290 meV. The self consistently determined results for various phy sical quantities are
given in Table III for magnetic order in the xdirection. All ordering directions within the
x−yplane are nearly equivalent. Besides the dominant Hund’s coupling indu ced easy-plane
anisotropy,53there is an extremely weak easy-axis anisotropy which will be discuss ed at the
end of this section. The octahedral rotation induces small in-plane canting of spins but the
canting axis is free to orient in any direction. The strong Coulomb inte raction induced SOC
renormalization by nearly 2/3 (Table IV) agrees with the pseudo-or bital based approach.51
TABLE III: Self consistently determined magnetization and density values, showing small spin
canting about the zaxis due to octahedral rotation induced DM interaction. Her etm1= 0.2.
µ(s)mx
µmy
µmz
µnµ
yz(A) 0.186 −0.052 0 1.653
xz(A)−0.185 0.049 0 1.654
xy(A)−0.172−0.047 0 1.693µ(s)mx
µmy
µmz
µnµ
yz(B)−0.186−0.052 0 1.653
xz(B) 0.185 0.049 0 1.654
xy(B) 0.172 −0.047 0 1.69320
TABLE IV: Self consistently determined renormalized SOC va luesλα=λ+λint
αand the orbital
magnetic moments /angbracketleftLα/angbracketrightforα=x,y,zon the two sublattices. Bare SOC value λ=1.35.
s λ xλyλz/angbracketleftLx/angbracketright /angbracketleftLy/angbracketright /angbracketleftLz/angbracketright
A 1.882 1.882 1.871 0.367 0.091 0
B 1.882 1.882 1.871 −0.367 0.091 0
The strong orbital moments and their correlation with the magnetic order direction (Table
IV) reflect the strong SOC induced spin-orbital entanglement.
The low energy part of the spectral function (Fig. 6(a)) clearly sh ows the gapless and
gapped modes corresponding to in-plane and out-of-plane fluctua tions, consistent with the
easy-plane anisotropy. The magnon gap ≈45 meV is close to the result obtained using the
pseudo-orbital based approach,53and in agreement with recent experiments.46It should be
noted that along with the full generalized spin sector, the orbital o ff-diagonal charge sector
(ψ†
µ1ψν)relatedtotheorbitalmomentoperators Lx,y,zwasincludedintheabovecalculations
which allows for the accompanying transverse fluctuations of orbit al moments. Indeed, the
exactly gapless Goldstone mode seen in Fig. 6(a) is obtained only if the ψ†
µ1ψνsector is
included, indicating the coupled spin-orbital nature of the Goldston e mode, as illustrated in
(π/2,π/2) (π,0) ( π,π) (0,0) 0 50 100 150 200 250 300ω (meV)
0.01 0.1 1 10 100
(a)
(π/2,π/2) (π,0) ( π,π) (0,0) 300 400 500 600 700 800ω (meV)
0.01 0.1 1 10 100
(b)
FIG. 6: The spectral function in the self-consistent state f or then= 5 case with planar AFM
order including octahedral rotation, showing the (a) gaple ss and gapped modes corresponding to
in-plane and out-of-plane fluctuations and (b) the spin-orb it exciton modes near 500 meV and 300
meV in the high-energy part.21
(π/2,π/2) (π,0) ( π,π) (0,0) 0 200 400 600 800 1000 1200ω (meV)
0.01 0.1 1 10 100
(a)JH = 0
(π/2,π/2) (π,0) ( π,π) (0,0) 0 0.4 0.8 1.2 1.6 2 2.4ω (eV)
0.01 0.1 1 10 100
(b)
λ = 5, JH = U/7
(π/2,π/2) (π,0) ( π,π) (0,0) 1.9 2 2.1 2.2 2.3ω (eV)
0.01 0.1 1 10 100
(c)
λ = 5, JH = U/7
FIG. 7: The spin-orbit exciton modes ( n= 5) for special cases showing (a) no splitting in the weak
branch (∼400 meV) for Hund’s coupling JH= 0, (b) disappearance of the weak branch for large
SOC value λ= 5, and (c) expanded view of the multiple exciton modes ( ∼2 eV) in case (b). Here
octahedral rotation is included in all three cases.
Appendix C showing the detailed spin-orbital composition.
Fig. 6(b) shows the spin-orbit exciton modes ( ∼500 meV) involving particle-hole excita-
tions between the J=1/2 and 3/2 sectors, which matches closely with results obtained u sing
the pseudo-orbital based approach.51As discussed in the previous ( n= 3) case, collective
modes arise from particle-hole excitations which are converted to w ell defined propagating
modes split off from the continuum by the Coulomb interaction induced resonant scattering
mechanism. The significantly weaker modes ( ∼300 meV) just below the particle-hole con-
tinuum for the nominally J= 1/2 sector are also spin-orbit exciton modes. The splitting
seen beyond ( π,0) vanishes for JH= 0, as seen in Fig. 7(a). The weak intensity corresponds
to the small J= 3/2 character (mainly mJ=±3/2) in the nominally J= 1/2 bands due
to strong mixing between the two sectors induced by the band (hop ping) terms. For large
SOC strength λ, the weak exciton modes disappear (Fig. 7(b)), confirming the abo ve pic-
ture. Thus, the (low) intensity of the weak exciton modes provides a direct measure of the
mixing between the J= 1/2 and 3/2 sectors. The fine splitting of exciton bands in Fig.
7(c) corresponds to four possible mJvalues (±3/2,±1/2) for the hole in the J= 3/2 sector
and the exciton hopping terms connecting the two sublattices.
We now discuss the extremely weak easy-axis anisotropy which leads to preferred isospin
(J= 1/2) orientation along the diagonal directions ±(ˆx±ˆy) within the easy plane. Fig.
8(a) shows the small magnon gap ( ≈3 meV) for the in-plane magnon mode induced by the
Hund’s coupling JHdue to the extremely weak spin twisting as shown in Fig. 8(b) which
results in an easy-axis anisotropy with C4symmetry. Here the parameter set is same as22
0 2 4 6 8 10 12 14
0 0.02 0.04 0.06 0.08in-plane mode(a)ωq (meV)
qx = qy
(b)
isospin
FIG. 8: (a) Magnon energies for isospin order along ˆ x+ˆydirection showing the small magnon gap
(≈3 meV) for the in-plane fluctuation mode. (b) The isospin and yz,xz,xy moment orientations
fortheideal spin-orbital entangled state, whichis extrem ely weakly perturbedbyfinite JHresulting
in slight twisting of the yz,xzmoments as indicated, leading to the easy-axis anisotropy w ithC4
symmetry. The isospin easy axes are along φ=nπ/4 wheren= 1,3,5,7.
earlier including the octahedral rotation which only weakly enhances the magnon gap. The
above weak perturbative effect of JHonthestrongly spin-orbital entangled statecorresponds
to the opposite end of the competition between SOC and JHas compared to the n= 3 case
discussed in Sec. V.
VII.n= 4— APPLICATION TO Ca2RuO4
For moderate tetragonal distortion ( ǫxy≈ −1), thexyorbital in the 4 d4compound
Ca2RuO4is nominally doubly occupied and magnetically inactive, while the nominally h alf-
filledandmagnetically active yz,xzorbitalsyieldaneffectively two-orbitalmagneticsystem.
Hund’s coupling between the two S= 1/2 spins results in low-lying (in-phase) and apprecia-
bly gapped (out-of-phase) spin fluctuation modes. The in-phase m odes of the yz,xzorbital
S= 1/2 spins correspond to an effective S= 1 spin system. However, the rich interplay be-
tween SOC, Coulomb interaction, octahedral rotations, and tetr agonal distortion results in
complex magnetic behaviour which crucially involves the xyorbital and is therefore beyond
the above simplistic picture.
Treating all the different physical elements on the same footing with in the unified frame-
work of the generalized self-consistent approach explicitly shows t he variety of physical23
(π/2,π/2) (π,0) ( π,π) (0,0) 0 20 40 60 80 100 120 140 160 180ω (meV)
0.01 0.1 1 10 100
(a)
(π/2,π/2) (π,0) ( π,π) (0,0) 250 300 350 400 450 500ω (meV)
0.01 0.1 1 10 100
(b)
FIG. 9: The generalized fluctuation spectral function for th en= 4 case, showing coupled spin-
orbital excitations including low-energy magnon modes (be low∼60 meV), intermediate-energy
orbiton (100 and 140 meV) and spin-orbiton (300 and 350 meV) m odes, and high-energy spin-
orbit exciton (425 meV) modes.
effects arising from the rich interplay in Ca 2RuO4. These include: SOC induced easy-plane
and easy-axis anisotropies similar to the n=3 case, octahedral tilting induced reduction of
easy-axis anisotropyfrom C4toC2symmetry, spin-orbital coupling induced orbital magnetic
moments, Coulomb interaction induced strongly anisotropic SOC ren ormalization, decreas-
ing tetragonal distortion induced magnetic reorientation transitio n from planar AFM order
to FM (z) order, and orbital moment interaction induced orbital gap.52Stable FM and
AFM metallic states were also obtained near the magnetic phase boun dary separating the
two magnetic orders. The self-consistent determination of magne tic order has also explicitly
shown the coupled nature of spin and orbital fluctuations, as refle cted in the ferro and an-
tiferro orbital fluctuations associated with in-phase and out-of- phase spin twisting modes,
highlighting the strong deviation from conventional Heisenberg beh aviour in effective spin
models, as discussed recently to account for the magnetic excitat ion measurements in INS
experiments on Ca 2RuO4.35
In the following, we will take the same parameter set as considered in the self-consistent
study,52along with U= 8 andJH=U/5 in the energy scale unit (150 meV), so that
U= 1.2 eV,U′′=U/2 = 0.6 eV, andJH= 0.24 eV. These are comparable to reported
values extracted from RIXS ( JH= 0.34 eV) and ARPES ( JH= 0.4 eV) studies.24,40The
hopping parameter values considered are as given in Sec. II, and th e bare SOC value λ= 1.24
Fig. 9 shows the calculated generalized fluctuation spectral funct ion. Several well de-
fined propagating modes are seen here including: (i) the low-energy (below∼60 meV)
dominantly spin (magnon) excitations involving the magnetically active yz,xzorbitals and
corresponding to in-plane and out-of-plane fluctuations which are gapped due to the mag-
netic anisotropies, (ii) the intermediate-energy (100 and 140 meV) dominantly orbital exci-
tations (orbitons) involving particle-hole excitations between xy(hole) andyz,xz(particle)
states, (iii) the intermediate-energy (300 and 350 meV) dominantly spin-orbital excitations
(spin-orbitons) involving xy(hole) and yz,xz(particle) states, and (iv) the high-energy
(425 meV) dominantly spin-orbital excitations (spin-orbit excitons ) involving particle-hole
excitations between yz,xz(hole) and yz,xz(particle) states of nominally different Jsec-
tors. The SOC-induced spin-orbital entangled Jstates are strongly renormalized by the
tetragonal splitting and the electronic correlation induced stagge red field.
The spin-orbital characterization of the various collective excitat ions mentioned above is
inferred from the basis-resolved contributions to the total spec tral functions which explicitly
show the relative spin-orbital composition of the various excitation s (Appendix C). The
presence of sharply defined collective excitations for the magnon, orbiton, and spin-orbiton
modes which are clearly separated from the particle-hole continuum highlights the rich spin-
orbital physics in the n= 4 case corresponding to the Ca 2RuO4compound. Many of our
calculated magnon spectra features such as the magnon gaps for in-plane and out-of-plane
modes, weak dispersive nature along the magnetic zone boundary, as well as the overall
magnon energy scale are in excellent agreement with the INS study.34,35The orbiton mode
energy scale is also qualitatively comparable to the composite excitat ion peaks obtained
around 80 meV in Raman and RIXS studies.37–40The calculated spin-orbiton and spin-orbit
exciton energies are also in agreement with the excitation peaks obt ained around 300-350
meV energy range and 400 meV in RIXS studies. We also obtained excit ations in the high-
energy range 750-800 meV and 900 meV (not shown), which are com parable to the peaks
obtained around 750 meV and 1000 meV in RIXS studies.25
VIII. CONCLUSIONS
Following up on the generalized self-consistent approach including or bital off-diagonal
spin and charge condensates, investigation of the generalized fluc tuation propagator reveals
thecomposite spin-orbital character ofthe different types ofco llective excitations instrongly
spin-orbit coupled systems. A realistic representation of magnetic anisotropy effects due to
the interplay of SOC, Coulomb interaction, and structural distort ion terms was included
in the three-orbital model, while maintaining uniformity of lattice stru cture in order to
focus on the coupled spin-orbital excitations. Our unified investiga tion of the three electron
filling cases n= 3,4,5 corresponding to the three compounds NaOsO 3, Ca2RuO4, Sr2IrO4
provides deep insight into how the spin-orbital physics in the magnet ic ground state is
reflected in the collective excitations. The calculated spectral fun ctions show well defined
propagating modes corresponding to dominantly spin (magnon), or bital (orbiton), and spin-
orbital (spin-orbiton) excitations, along with the spin-orbit excito n modes involving spin-
orbital excitations between states of different Jsectors induced by the spin-orbit coupling.
Appendix A: Orbital off-diagonal condensates in the HF approximation
TheadditionalcontributionsintheHFapproximationarisingfromthe orbitaloff-diagonal
spin and charge condensates are given below. For the density, Hun d’s coupling, and pair
hopping interaction terms in Eq. 3, we obtain (for site i):
U′′/summationdisplay
µ<νnµnν→ −U′′
2/summationdisplay
µ<ν[nµν/angbracketleftnνµ/angbracketright+σµν./angbracketleftσνµ/angbracketright]+H.c.
−2JH/summationdisplay
µ<νSµ.Sν→JH
4/summationdisplay
µ<ν[3nµν/angbracketleftnνµ/angbracketright−σµν./angbracketleftσνµ/angbracketright]+H.c.
JP/summationdisplay
µ/negationslash=νa†
µ↑a†
µ↓aν↓aν↑→JP
2/summationdisplay
µ<ν[nµν/angbracketleftnµν/angbracketright−σµν./angbracketleftσµν/angbracketright]+H.c. (A1)
in terms of the orbital off-diagonal spin ( σµν=ψ†
µσψν) and charge ( nµν=ψ†
µ1ψν) oper-
ators. The orbital off-diagonal condensates are finite due to the SOC-induced spin-orbital
correlations. These additional terms in the HF theory explicitly pres erve the SU(2) spin
rotation symmetry of the various Coulomb interaction terms.
Collecting all the spin and charge terms together, we obtain the orb ital off-diagonal26
(OOD) contributions of the Coulomb interaction terms:
[HHF
int]OOD=/summationdisplay
µ<ν/bracketleftbigg/parenleftbigg
−U′′
2+3JH
4/parenrightbigg
nµν/angbracketleftnνµ/angbracketright+/parenleftbiggJP
2/parenrightbigg
nµν/angbracketleftnµν/angbracketright
−/parenleftbiggU′′
2+JH
4/parenrightbigg
σµν./angbracketleftσνµ/angbracketright−/parenleftbiggJP
2/parenrightbigg
σµν./angbracketleftσµν/angbracketright/bracketrightbigg
+H.c. (A2)
Appendix B: Coulomb interaction matrix elements in the orbital-pair basis
CorrespondingtotheaboveHFcontributionsintheorbitaloff-diag onalsector, weexpress
the Coulomb interactions in terms of the generalized spin and charge operators (for site i):
[Hint]OOD=/summationdisplay
µ<ν/bracketleftbigg/parenleftbigg
−U′′
2+3JH
4/parenrightbigg
nµνn†
µν−/parenleftbiggU′′
2+JH
4/parenrightbigg
σµν.σ†
µν/bracketrightbigg
+/summationdisplay
µ<ν/bracketleftbigg/parenleftbiggJP
4/parenrightbigg
nµνn†
νµ−/parenleftbiggJP
4/parenrightbigg
σµν.σ†
νµ+H.c./bracketrightbigg
(B1)
wheren†
µν=nνµandσ†
µν=σνµ. The above form shows that only the pair-hopping
interaction terms ( JP) are off-diagonal in the orbital-pair ( µν) basis. We will use the above
Coulomb interaction terms in the orbital off-diagonal sector in the R PA series in order to
ensure consistency with the self-consistent determination of mag netic order including the
orbital off-diagonal condensates.
The Coulomb interaction terms in the orbital diagonal sector can be cast in a similar
form:
[Hint]OD=/summationdisplay
µ/bracketleftbigg/parenleftbigg
−U
4/parenrightbigg
σµ.σµ+/parenleftbiggU
4/parenrightbigg
nµnµ/bracketrightbigg
+/summationdisplay
µ<ν/bracketleftbigg/parenleftbigg
−2JH
4/parenrightbigg
σµ.σν+U′′nµnν/bracketrightbigg
(B2)
which include the Hubbard, Hund’s coupling, and density interaction t erms.
The form of the [ U] matrix used in the RPA series Eq. (17) is now discussed below. In
the composite spin-charge-orbital-sublattice ( µναs) basis, the [ U] matrix is diagonal in spin,
charge, and sublattice sectors. There are two possible cases invo lving the orbital-pair ( µν)
basis. In the case µ=ν, the [U] matrices in the spin ( α=x,y,z) and charge ( α=c) sectors
are obtained as:
[U]µ′µ′α′=α
µµα=x,y,z=
U JHJH
JHU JH
JHJHU
[U]µ′µ′α′=α
µµα=c=
−U−2U′′−2U′′
−2U′′−U−2U′′
−2U′′−2U′′−U
(B3)27
FIG. 10: The basis-resolved contributions to the total spec tral function for the low-energy magnon
(left panel) and intermediate-energy orbiton (center and r ight panels) modes, showing dominantly
spin (µ=ν,α=x,y,z) and orbital ( µ/negationslash=ν,α=c) character of the fluctuation modes, respectively.
corresponding to the interaction terms (Eq. B2) for the normal s pin and charge density
operators. Similarly, for the six orbital-pair cases ( µ,ν) corresponding to µ/negationslash=ν, the [U]
matrix elements in the spin ( α=x,y,z) and charge ( α=c) sectors are obtained as:
[U]µνα
µνα=x,y,z=U′′+JH/2 [U]µνα
µνα=c=U′′−3JH/2
[U]νµα
µνα=x,y,z=JP [U]νµα
µνα=c=−JP (B4)
corresponding to the interaction terms (Eq. B1) involving the orbit al off-diagonal spin and
charge operators.
Appendix C: Basis-resolved contributions to the total spectral function
The detailed spin-orbital character of the collective excitations ca n be identified from the
basis-resolved contributions to the total spectral functions. T his is illustrated here for the28
FIG. 11: The basis-resolved contributions to the total spec tral function for the intermediate-
energy spin-orbiton (left and center panels) and high-ener gy spin-orbit exciton (right panel) modes,
showing dominantly spin-orbital character ( µ/negationslash=ν,α=x,y,z) involving xyandyz,xzorbitals (left
and center panels) and yz,xzorbitals (right panel).
excitations shown in Fig. 9 for the n= 4 case corresponding to the Ca 2RuO4compound.
Fig. 10 shows dominantly spin excitations involving yz,xzorbitals for the magnon modes
(below 60 meV) and dominantly orbital excitations involving xyandyz,xzorbitals for
the orbiton modes (100 and 140 meV). Similarly, Fig. 11 shows dominan tly spin-orbital
excitations involving xyandyz,xzorbitals for the spin-orbiton modes (300 and 350 meV),
and dominantly spin-orbital excitations involving yz,xzorbitals for the spin-orbit exciton
modes (425 meV).29
FIG. 12: The extreme spin-orbital-entanglement induced co rrespondence between (a) magnetic
ordering directions, (b) sign of magnetic moments for the th ree orbitals, and (c) orbital current
induced orbital moments for the three orbitals, for the n= 5 case corresponding to Sr 2IrO4.
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
x y z cω = 0 meV
(a)π-1Im[χ(q,ω)]µναµναµν=yz yz
xz xz
xy xy
yz xz
xz xyxy yz
xz yz
xy xz
yz xyq =(0,0)
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
x y z cω = 46 meV
(b)π-1Im[χ(q,ω)]µναµναq =(0,0)
FIG. 13: The basis-resolved contributions to the total spec tral function for the (a) gapless in-plane
magnon mode and (b) gapped out-of-plane magnon mode for the n= 5 case corresponding to
Sr2IrO4with extreme spin-orbital entanglement.
Similarly, for the n= 5 case corresponding to Sr 2IrO4, the detailed spin-orbital character
of the Goldstone mode and gapped mode at q= (0,0) seen in Fig. 6 is shown in Fig.
13, explicitly illustrating the effect of extreme spin-orbital entangle ment and the resulting
correspondence (Fig. 12) between magnetic ordering directions, spin moments, and orbital
moments.30
∗Electronic address: avinas@iitk.ac.in
1W. Witczak-Krempa, G. Chen, Y. B. Kim, and L. Balents, Correlated Quantum Phenomena in
the Strong Spin-Orbit Regime , Annu. Rev. Condens. Matter Phys. 5, 57 (2014).
2J. G. Rau, E. Kin-Ho Lee, and H.-Y. Kee, Spin-Orbit Physics Giving Rise to Novel Phases in
Correlated Systems: Iridates and Related Materials , Annu. Rev. Condens. Matter Phys. 7, 195
(2016).
3G. Cao and P. Schlottmann, The Challenge of Spin-Orbit-Tuned Ground States in Iridates: A
Key Issues Review , Rep. Prog. Phys. 81, 042502 (2018).
4G. Zhang and E. Pavarini, Spin-Orbit and Coulomb Effects in Single-Layered Ruthenates , Phys.
Status Solidi RRL 12, 1800211 (2018).
5J. Bertinshaw, Y. K. Kim, G. Khaliullin, and B. J. Kim, Square Lattice Iridates , Annu. Rev.
Condens. Matter Phys. 10, 315 (2019).
6G. Zhang and E. Pavarini, Higgs Mode and Stability of xy-orbital Ordering in Ca2RuO4, Phys.
Rev. B101, 205128 (2020).
7Y. G. Shi, Y. F. Guo, S. Yu, M. Arai, A. A. Belik, A. Sato, K. Yama ura, E. Takayama-
Muromachi, H. F. Tian, H. X. Yang, J. Q. Li, T. Varga, J. F. Mitc hell, and S. Okamoto,
Continuous Metal-Insulator Transition of the Antiferromagn etic Perovskite NaOsO 3, Phys.Rev.
B80, 161104(R) (2009).
8S. Calder, V. O. Garlea, D. F. McMorrow, M. D. Lumsden, M. B. St one, J. C. Lang, J.-W.
Kim, J. A. Schlueter, Y. G. Shi, K. Yamaura, Y. S. Sun, Y. Tsuji moto, and A. D. Christian-
son,Magnetically Driven Metal-Insulator Transition in NaOsO 3, Phys. Rev. Lett. 108, 257209
(2012).
9Y. Du, X. Wan, L. Sheng, J. Dong, and S. Y. Savrasov, Electronic Structure and Magnetic
Properties of NaOsO 3, Phys. Rev. B 85, 174424 (2012).
10M.-C. Jung, Y.-J. Song, K.-W. Lee, and W. E. Pickett, Structural and Correlation effects in the
Itinerant Insulating Antiferromagnetic Perovskite NaOsO 3, Phys. Rev. B 87, 115119 (2013).
11I. Lo Vecchio, A. Perucchi, P. Di Pietro, O. Limaj, U. Schade, Y. Sun, M. Arai, K. Yamaura,
and S. Lupi, Infrared evidence of a Slater Metal-Insulator Transition in NaOsO 3, Sci. Rep. 3,
2990 (2013).31
12B. Kim, P. Liu, Z. Erg¨ onenc, A. Toschi, S. Khmelevskyi, and C . Franchini, Lifshitz Transition
Driven by Spin Fluctuations and Spin-Orbit Renormalizatio n inNaOsO 3, Phys. Rev. B 94,
241113(R) (2016).
13J.G. Vale, S.Calder, C.Donnerer, D. Pincini, Y. G. Shi, Y. Ts ujimoto, K. Yamaura, M. Moretti
Sala, J. van den Brink, A. D. Christianson, and D. F. McMorrow ,Evolution of the Magnetic
Excitations in NaOsO 3through its Metal-Insulator Transition , Phys. Rev. Lett. 120, 227203
(2018).
14S. Mohapatra, C. Bhandari, S. Satpathy, and A. Singh, Effect of Structural Distortion on the
Electronic Band Structure of NaOsO 3studied within Density Functional Theory and a Three-
Orbital Model , Phys. Rev. B 97, 155154 (2018).
15A. Singh, S. Mohapatra, C. Bhandari, and S. Satpathy, Spin-Orbit Coupling Induced Magnetic
Anisotropy and Large Spin Wave Gap in NaOsO 3, J. Phys. Commun. 2, 115016 (2018).
16S. Calder, J. G. Vale, N. Bogdanov, C. Donnerer, D. Pincini, M . Moretti Sala, X. Liu, M. H.
Upton, D. Casa, Y. G. Shi, Y. Tsujimoto, K. Yamaura, J. P. Hill , J. van den Brink, D. F.
McMorrow, and A. D. Christianson, Strongly Gapped Spin-Wave Excitation in the Insulating
Phase of NaOsO 3, Phys. Rev. B 95, 020413(R) (2017).
17J.G. Vale, S.Calder, C.Donnerer, D. Pincini, Y. G. Shi, Y. Ts ujimoto, K. Yamaura, M. Moretti
Sala, J. van den Brink, A. D. Christianson, and D. F. McMorrow ,Crossover from Itinerant to
Localized Magnetic Excitations Through the Metal-Insulator Transition in NaOsO 3, Phys. Rev.
B97, 184429 (2018).
18S. Nakatsuji, S.-i. Ikeda, and Y. Maeno, Ca 2RuO4: New Mott Insulators of Layered Ruthenate ,
J. Phys. Soc. Jpn 66(7), 1868 (1997).
19M.Braden, G.Andr´ e, S.Nakatsuji, andY. Maeno, Crystal and Magnetic Structure of Ca2RuO4:
Magnetoelastic Coupling and the Metal-Insulator Transition , Phys. Rev. B 58, 847 (1998).
20C. S. Alexander, G. Cao, V. Dobrosavljevic, S. McCall, J. E. C row, E. Lochner, and R. P.
Guertin, Destruction of the Mott Insulating Ground State of Ca2RuO4by a Structural Transi-
tion, Phys. Rev. B 60, R8422 (1999).
21O. Friedt, M. Braden, G. Andr´ e, P. Adelmann, S. Nakatsuji, a nd Y. Maeno, Structural and
Magnetic Aspects of the Metal-Insulator Transition in Ca2−xSrxRuO4, Phys. Rev. B 63, 174432
(2001).
22E. Gorelov, M. Karolak, T. O. Wehling, F. Lechermann, A. I. Li chtenstein, and E. Pavarini,32
Nature of the Mott Transition in Ca2RuO4, Phys. Rev. Lett. 104, 226401 (2010).
23G. Zhang and E. Pavarini, Mott transition, Spin-Orbit Effects, and Magnetism in Ca2RuO4,
Phys. Rev. B 95, 075145 (2017).
24D. Sutter, C. G. Fatuzzo, S. Moser, M. Kim, R. Fittipaldi, A. V ecchione, V. Granata, Y. Sassa,
F. Cossalter, G. Gatti, M. Grioni, H. M. Rønnow, N. C. Plumb, C . E. Matt, M. Shi, M. Hoesch,
T. K. Kim, T.-R. Chang, H.-T. Jeng, C. Jozwiak, A. Bostwick, E . Rotenberg, A. Georges, T.
Neupert, and J. Chang, Hallmarks of Hund’s Coupling in the Mott Insulator Ca2RuO4, Nat.
Comm.8, 15176 (2017).
25F. Nakamura, T. Goko, M. Ito, T. Fujita, S. Nakatsuji, H. Fuka zawa, Y. Maeno, P. Alireza, D.
Forsythe, and S. R. Julian, From Mott Insulator to Ferromagnetic Metal: A Pressure Study of
Ca2RuO4, Phys. Rev. B 65, 220402(R) (2002).
26P. Steffens, O. Friedt, P. Alireza, W. G. Marshall, W. Schmidt, F. Nakamura, S. Nakatsuji, Y.
Maeno, R. Lengsdorf, M. M. Abd-Elmeguid, and M. Braden, High-pressure Diffraction Studies
onCa2RuO4, Phys. Rev. B 72, 094104 (2005).
27H. Taniguchi, K. Nishimura, R. Ishikawa, S. Yonezawa, S. K. G oh, F. Nakamura, and Y.
Maeno,Anisotropic Uniaxial Pressure Response of the Mott Insulato rCa2RuO4, Phys. Rev. B
88, 205111 (2013).
28S. Nakatsuji and Y. Maeno, Quasi-Two-Dimensional Mott Transition System Ca2−xSrxRuO4,
Phys. Rev. Lett. 84, 2666 (2000).
29Z. Fang and K. Terakura. Magnetic Phase Diagram of Ca2−xSrxRuO4Governed by Structural
Distortions , Phys. Rev. B 64, 020509 (2001).
30P. Steffens, O. Friedt, Y. Sidis, P. Link, J. Kulda, K. Schmalzl , S. Nakatsuji, and M. Braden,
Magnetic Excitations in the Metallic Single-Layer Ruthena tesCa2−xSrxRuO4Studied by Inelas-
tic Neutron Scattering , Phys. Rev. B 83, 054429 (2011).
31C. Dietl, S. K. Sinha, G. Christiani, Y. Khaydukov, T. Keller , D. Putzky, S. Ibrahimkutty, P.
Wochner, G. Logvenov, P. A. van Aken, B. J. Kim, and B. Keimer, Tailoring the Electronic
Properties of Ca2RuO4via Epitaxial Strain , Appl. Phys. Lett. 112, 031902 (2018).
32F. Nakamura, M. Sakaki, Y. Yamanaka, S. Tamaru, T. Suzuki, an d Y. Maeno, Electric-Field-
Induced Metal Maintained by Current of the Mott Insulator Ca2RuO4, Sci. Rep. 3, 2536 (2013).
33R. Okazaki, Y. Nishina, Y. Yasui, F. Nakamura, T. Suzuki, and I. Terasaki, Current-Induced
Gap Suppression in the Mott Insulator Ca2RuO4, J. Phys. Soc. Jpn. 82, 103702 (2013).33
34S. Kunkem¨ oller, D. Khomskii, P. Steffens, A. Piovano, A. A. Nu groho, and M. Braden, Highly
Anisotropic Magnon Dispersion in Ca2RuO4: Evidence for Strong Spin-Orbit Coupling , Phys.
Rev. Lett. 115, 247201 (2015).
35A. Jain, M. Krautloher, J. Porras, G. H. Ryu, D. P. Chen, D. L. A bernathy, J. T. Park, A.
Ivanov, J. Chaloupka, G. Khaliullin, B. Keimer, and B. J. Kim ,Higgs Mode and Its Decay in a
Two-Dimensional Antiferromagnet , Nat. Phys. 13, 633 (2017).
36S. Kunkem¨ oller, E. Komleva, S. V. Streltsov, S. Hoffmann, D. I . Khomskii, P. Steffens, Y. Sidis,
K. Schmalzl, and M. Braden, Magnon dispersion in Ca2Ru1−xTixO4: Impact of Spin-Orbit
Coupling and Oxygen Moments , Phys. Rev. B 95, 214408 (2017).
37S.-M. Souliou, J. Chaloupka, G. Khaliullin, G. Ryu, A. Jain, B. J. Kim, M. Le Tacon, and B.
Keimer, Raman Scattering from Higgs Mode Oscillations in the Two-Dim ensional Antiferro-
magnetCa2RuO4, Phys. Rev. Lett. 119, 067201 (2017).
38C. G. Fatuzzo, M. Dantz, S. Fatale, P. Olalde-Velasco, N. E. S haik, B. Dalla Piazza, S. Toth,
J. Pelliciari, R. Fittipaldi, A. Vecchione, N. Kikugawa, J. S. Brooks, H. M. Rønnow, M. Grioni,
Ch. R¨ uegg, T. Schmitt, and J. Chang, Spin-Orbit-Induced Orbital Excitations in Sr2RuO4and
Ca2RuO4: A resonant Inelastic X-Ray Scattering Study , Phys. Rev. B 91, 155104 (2015).
39L. Das, F. Forte, R. Fittipaldi, C. G. Fatuzzo, V. Granata, O. Ivashko, M. Horio, F. Schindler,
M. Dantz, Yi Tseng, D. E. McNally, H. M. Rønnow, W. Wan, N. B. Ch ristensen, J. Pelliciari, P.
Olalde-Velasco, N. Kikugawa, T. Neupert, A. Vecchione, T. S chmitt, M. Cuoco, and J. Chang,
Spin-Orbital Excitations in Ca2RuO4Revealed by Resonant Inelastic X-Ray Scattering , Phys.
Rev. X8, 011048 (2018).
40H. Gretarsson, H. Suzuki, Hoon Kim, K. Ueda, M. Krautloher, B . J. Kim, H. Yava¸ s, G. Khali-
ullin, and B. Keimer, Observation of Spin-Orbit Excitations and Hund’s Multiple ts inCa2RuO4,
Phys. Rev. B 100, 045123 (2019).
41B. J. Kim, H. Jin, S. J. Moon, J.-Y. Kim, B.-G. Park, C. S. Leem, J. Yu, T. W. Noh, C.
Kim, S.-J. Oh, J.-H. Park, V. Durairaj, G. Cao and E. Rotenber g,NovelJeff= 1/2Mott State
Induced by Relativistic Spin-Orbit Coupling in Sr2IrO4, Phys. Rev. Lett. 101, 076402 (2008).
42B.J.Kim, H.Ohsumi,T.Komesu, S.Sakai, T.Morita, H. Takagi andT.Arima, Phase-Sensitive
Observation of a Spin-Orbital Mott State in Sr2IrO4, Science 323, 1329–1332 (2009).
43J. Kim, D. Casa, M. H. Upton, T. Gog, Y.-J. Kim, J. F. Mitchell, M. van Veenendaal, M.
Daghofer, J.vandenBrink, G.Khaliullin, andB.J.Kim, Magnetic Excitation Spectra of Sr2IrO434
Probed by Resonant Inelastic X-Ray Scattering: Establishin g Links to Cuprate Superconductors ,
Phys. Rev. Lett. 108, 177003 (2012).
44J-i Igarashi and T. Nagao, Interplay between Hund’s Coupling and Spin-Orbit Interaction on
Elementary Excitations in Sr2IrO4, J. Phys. Soc. Jpn. 83, 053709 (2014).
45D. Pincini, J. G. Vale, C. Donnerer, A. de la Torre, E. C. Hunte r, R. Perry, M. Moretti Sala,
F. Baumberger, and D. F. McMorrow, Anisotropic Exchange and Spin-Wave Damping in Pure
and Electron-Doped Sr2IrO4, Phys. Rev. B 96, 075162 (2017).
46J. Porras, J. Bertinshaw, H. Liu, G. Khaliullin, N. H. Sung, J .-W. Kim, S. Francoual, P.
Steffens, G. Deng, M. Moretti Sala, A. Effimenko, A. Said, D. Casa , X. Huang, T. Gog, J.
Kim, B. Keimer, and B. J. Kim, Pseudospin-Lattice Coupling in the Spin-Orbit Mott Insulato r
Sr2IrO4, Phys. Rev. B 99, 085125 (2019).
47S. Mohapatra, J. van den Brink, and A. Singh, Magnetic Excitations in a Three-Orbital Model
for the Strongly Spin-Orbit Coupled Iridates: Effect of Mixing Between the J= 1/2and3/2
Sectors, Phys. Rev. B 95, 094435 (2017).
48J. Kim, M. Daghofer, A. H. Said, T. Gog, J. van den Brink, G. Kha liullin, and B. J. Kim,
Excitonic Quasiparticles in a Spin-Orbit Mott Insulator , Nat. Commun. 5, 4453 (2014).
49O. Krupin, G. L. Dakovski, B. J. Kim, J. Kim, S. Mishra, Yi-De C huang, C. R. Serrao, W.-S.
Lee, W. F. Schlotter, M. P. Minitti, D. Zhu, D. Fritz, M. Choll et, R. Ramesh, S. L. Molodtsov,
and J. J. Turner, Ultrafast Dynamics of Localized Magnetic Moments in the Unc onventional
Mott Insulator Sr2IrO4, J. Phys.: Condens. Matter 28, 32LT01 (2016).
50M. Souri, B. H. Kim, J. H. Gruenewald, J. G. Connell, J. Thomps on, J. Nichols, J. Terzic, B. I.
Min, G. Cao, J. W. Brill, and A. Seo, Optical Signatures of Spin-Orbit Exciton in Bandwidth-
Controlled Sr2IrO4Epitaxial Films via High-Concentration Ca and Ba Doping , Phys. Rev. B
95, 235125 (2017).
51S. Mohapatra and A. Singh, Correlated Motion of Particle-Hole Excitations Across the R enor-
malized Spin-Orbit Gap in Sr2IrO4, J. Magn. Magn. Mater 512, 166997 (2020).
52S. Mohapatra and A. Singh, Magnetic Reorientation Transition in a Three Orbital Model f or
Ca2RuO4— Interplay of Spin-Orbit Coupling, Tetragonal Distortion, a nd Coulomb Interac-
tions, J. Phys.: Condens. Matter 32, 485805 (2020).
53S. Mohapatra and A. Singh, Pseudo-Spin Rotation Symmetry Breaking by Coulomb Interacti on
terms in Spin-Orbit Coupled Systems , J. Phys.: Condens. Matter 33, 065802 (2021).35
54G. Khaliullin, Excitonic Magnetism in Van Vleck–type d4Mott Insulators , Phys. Rev. Lett.
111, 197201 (2013).
55A. Akbari and G. Khaliullin, Magnetic Excitations in a Spin-Orbit-Coupled d4Mott Insulator
on the Square Lattice , Phys. Rev. B 90, 035137 (2014).
56T. Feldmaier, P. Strobel, M. Schmid, P. Hansmann, and M. Dagh ofer,Excitonic Magnetism
at the Intersection of Spin-Orbit Coupling and Crystal-Field S plitting, Phys. Rev. Research 2,
033201 (2020).
57S.MohapatraandA.Singh, Spin Waves and Stability of Zigzag Order in the Hubbard Model with
Spin-Dependent Hopping Terms: Application to the Honeycom b Lattice Compounds Na2IrO3
andα-RuCl3, J. Magn. Magn. Mater 479, 229 (2019).(π/2,π/2) (π,0) ( π,π) (0,0) 600 650 700 750 800 850 900 950 1000ω (meV)
0.01 0.1 1 10 100
(c) |
1205.6629v1.Conservation_law_in_noncommutative_geometry____Application_to_spin_orbit_coupled_systems.pdf | arXiv:1205.6629v1 [math-ph] 30 May 2012Conservation law in noncommutative geometry
– Application to spin-orbit coupled systems
Naoyuki Sugimoto1and Naoto Nagaosa1,2
1Cross-correlated Materials Research Group (CMRG) and Corr elated
Electron Research Group (CERG), RIKEN, Saitama 351-0198, J apan
2Department of Applied Physics University of Tokyo, Tokyo 11 3-8656, Japan
The quantization scheme by noncommutative geometry develo ped in string theory is applied to
establish the conservation law of twisted spin and spin curr ent densities in the spin-orbit coupled
systems. Starting from the pedagogical introduction to Hop f algebra and deformation quantization,
the detailed derivation of the conservation law is given.2
CONTENTS
I. Introduction 3
II. Noether’s theorem in field theory 4
A. Conventional formulation of Noether’s theorem 4
B. Generalization of Noether’s theorem 7
III. Hopf algebra 8
1. Algebra 8
2. Coalgebra 10
3. Dual-algebra and Hopf algebra 11
IV. Deformation quantization 12
A. Wigner representation 13
B. Star product 14
1. Cohomology equation 16
2.L∞algebra 18
C. Topological string theory 20
1. Ghost fields and anti-fields 21
2. Condition of gauge invariance of classical action 22
3. Gauge invariance in path integral 23
D. Equivalence between deformation quantization and topological s tring theory 25
1. Path integral as L∞map 25
2. Perturbation theory 26
E. Diagram rules of deformation quantization 27
F. Gauge invariant star product 29
V. Twisted spin 31
A. Derivation of a twisted spin in Wigner space 31
B. Rashba-Dresselhaus model 33
VI. Conclusions 35
References 363
I. INTRODUCTION
Electrons are described by the Dirac equation where the U(1) Maxw ell electromagnetic field (emf) Aµis coupled
to the charge current jµas described by the Lagrangian (in the natural unit where /planckover2pi1=c= 1;µ= 0,1,2,3) [1]
L=¯ψ[iγµˆDµ−m]ψ. (1)
whereˆDµ=∂µ−ieAµis the covariant derivative, mis the electron mass. Note that the spin is encoded by 4
component nature of the spinors ψand¯ψ=ψ†γ0and the 4×4 gamma matrices γµ, but the charge and charge
current alone determine the electromagnetic properties of the ele ctrons, which are given by
jµ=−∂L
∂Aµ=−e¯ψγµψ. (2)
In condensed matter physics, on the other hand, the low energy p henomena compared with the mass gap 2 mc2∼
106eVare considered, and only the positive energy states described by t he two-component spinor are relevant. Then,
the relativistic spin-orbit interaction originates when the negative e nergy states (positron stats) are projected out
to derive the effective Hamiltonian or Lagrangian. The projection to a subspace of the Hilbert space leads to the
nontrivial geometrical structure which is often described by the g auge theory. This is also the case for the Dirac
equation, and the resultant gauge field is SU(2) non-Abelian gauge fi eld corresponding to the Zeeman effect (time-
component) and the spin-orbit interaction (spatial components) as described below.
The effective Lagrangianforthe positiveenergystatescan be der ivedby the expansionwith respect to 1 /(mc2) [2–4]
L= iψ†D0ψ+ψ†D2
2mψ+1
2mψ†/bracketleftbigg
eqσaA·Aa+q2
4Aa·Aa/bracketrightbigg
ψ, (3)
whereψis now the two-component spinor and D0=∂0+ieA0+iqAa
0σa
2, andDi=∂i−ieAi−iqAa
iσa
2(i= 1,2,3)
are the gauge covariant derivatives with qbeing the quantity proportional to the Bohr magneton [2, 4]. Aµis the
Maxwell emf, and the SU(2) gauge potential are defined as
Aa
0=Ba
Aa
i=ǫiaℓEℓ, (4)
andσx,y,zrepresent the Pauli matrices. The SU(2) gauge field is coupled to th e 4-component spin current
ja
0=ψσaψ,
ja
i=1
2mi[ψ†σaDiψ−Diψ†σaψ]. (5)
Namely, the Zeeman coupling and the spin-orbit interaction can be re garded as the gauge coupling between the 4-spin
current and the SU(2) gauge potential. (The spin current is the te nsor quantity with one suffix for the direction
of the spin polarization while the other for the direction of the flow.) N ote that the system has no SU(2) gauge
symmetry since the “vector potential” Aa
µis given by the physical field strength BandE, i.e., the relation ∂µAa
µ= 0
automatically holds. This fact is connected to the absence of the co nservation law for the spin density and spin
current density in the presence of the relativistic spin-orbit intera ction. In the spherically symmetric systems, the
total angular momentum, i.e., the sum of the orbital and spin angular momenta, is conserved, but the rotational
symmetry is usually broken by the periodic or disorder potential A0in condensed matter systems. Therefore, it is
usually assumed that the conservation law of spin is lost by the spin-o rbit interaction.
However, it is noted that the spin and spin current densities are “co variantly” conserved as described by the
“continuity equation” [2–4]
D0Ja
0+D·Ja= 0. (6)
replacing the usual derivative ∂µby the covariant derivative Dµ. This suggest that the conservation law holds in
the co-moving frame, but the crucial issue is how to translate this la w to the laboratory frame, which is the issue
addressed in this paper. Note again that the SU(2) gauge symmetr y is absent in the present problem, and hence the
Lagrangian like tr( FµνFµν), which usually leads to the generalized Maxwell equation and also to t he conservation
law of 4 spin current including both the matter field and gauge field [1], is missing. Instead, we will regard Aa
µas the
frozen background gauge field, and focus on the quantum dynamic s of noninteracting electrons only.4
In this paper, we derive the hidden conservation law by defining the “ twisted” spin and spin current densities which
satisfy the continuity equation with the usual derivative ∂µ. The description is intended to be pedagogical and self-
contained. For this purpose, the theoretical techniques develop ed in high energy physics is useful. The essential idea
is to take into account the effect of the background gauge field in te rms of the noncommutative geometry generalizing
the concept of “product”. This is achieved by extending the usual Lie algebra to Hopf algebra.
Usually, a conservation law is derived from symmetry of an action, i.e., Noether’s theorem. The symmetry in the
noncommutative geometry is called as a “twisted” symmetry, and th is symmetry and the corresponding generalized
Noether’s theorem have been studied in the high energy physics. Se iberg and Witten proposed that an equivalence of
a certain string theory and a certain field theory in noncommutative geometry [5]. Since then, the noncommutative
geometry have been attracted many researchers. On the other hand, it is known that the Poincaresymmetry is broken
in a field theory on a noncommutative geometry. It is a serious proble m because the energy and momentum cannot
be defined. M. Chaichian, et al. proposed the twisted symmetry in the Minkowski spacetime, and a lleged that the
twisted Poincare symmetry is substituted for the Poincare symmet ry [6, 7]. Moreover, G. Amelino-Camelia, et al.
discussed Noether’s theorem in the noncommutative geometry [8, 9 ].
As we will discuss in detail later, a certain type ofa noncommutative g eometry space is equal to a spin-orbit coupled
system. Therefore, a global SU(2) gauge symmetry in the noncom mutative geometry space gives a Noether current
corresponding to the “twisted” spin and spin current in the spin-or bit coupling system. This enables us to derive the
generalized Noether’s theorem for the twisted spin and spin curren t densities.
Now some remarks about the application is in order. Spintronics is an e merging field of electronics where the role
of charge and charge current are replaced by the spin and spin cur rent aiming at the low energy cost functions [10].
The relativistic spin-orbit interaction plays the key role there since it enables the manipulation of spins by the electric
field. However, this very spin-orbit interaction introduces the spin relaxation which destroys the spin information in
sharp contrast to the case of charge where the information is pro tected by the conservation law. Therefore, it has
been believed that the spintronics is possible in a short time-scale or t he small size devices. The discovery of the
conservation law of twisted spin and spin current densities means th at the quantum information of spin is preserved
by this hidden conservation law, and could be recovered. Actually, it has been recently predicted that the adiabatic
change in the spin-orbit interaction leads to the recovery of the sp in moment called spin-orbit echo [11]. Therefore,
the conservation law of the twisted spin and spin current densities is directly related to the applications in spintronics.
The plan ofthis paper follows (see Fig. 1). In section II, we reviewth e conventionalNoether’s theorem, and describe
briefly its generalization to motivate the use of Hopf algebra and def ormation quantization. In section III, the Hopf
algebra is introduced, and section IV gives the explanation of the de formation quantization with the star product.
The gauge interaction is compactly taken into account in the definitio n of the star product. These two sections are
sort of short review for the self-containedness and do not conta in any original results except the derivation of the star
product with gauge interaction. Section V is the main body of this pap er. By combining the Hopf algebra and the
deformation quantization, we present the derivation of the conse rved twisted spin and spin current densities. Section
VI is a brief summary of the paper and contains the possible new direc tions for future studies. The readers familiar
with the noncommutative geometry and deformation quantization c an skip sections III, IV, and directly go to section
V.
II. NOETHER’S THEOREM IN FIELD THEORY
In this section, we discuss Noether’s theorem [12], and its generaliza tion as a motivation to introduce the Hopf
algebra and deformation quantization. In section IIA, we will recall Noether theorem, and rewrite it using the so-
called “coproduct”, which is an element of the Hopf algebra. In sect ion IIB, we will sketch a derivation of generalized
Noether theorem.
A. Conventional formulation of Noether’s theorem
We start with the action Igiven by
I=/integraldisplay
ΩdDimxL(x)
=/integraldisplay
dDimxhΩ(x)L(x), (7)5
FIG. 1. Flows of derivation of generalized Noether’s theore m. Roman numerals and capital letters in boxes represent
section and subsection numbers, respectively. A generaliz ation of the Noether’s theorem is achieved through Hopf alge bra and
deformation quantization (section V). Hopf algebra appear to characterize feature of an infinitesimal transformed var iation
operator (sections II and III). The SU(2) gauge structure is embedded in the star product (section IV).
where Ω represents a range of the spacetime coordinate x(≡(x0,xi)≡(ct,x)) with a dimension Dim, i.e., (Dim −1)
is the dimension of the space, Ldescribes a Lagrangian density, and
hΩ(x) =/braceleftbigg
1 for{x|x∈Ω}
0 for{x|x /∈Ω}.; (8)
crepresents light speed. We introduce a field φrwith internal degree of freedom r, and infinitesimal transformations:
xµ/ma√sto→(x′)µ:=xµ+δζxµ, (9)
φr(x)/ma√sto→φ′
r(x′) :=φr(x)+δζφr(x), (10)
where we characterize the transformations by the subscript; sp ecifically,ζrepresents a general infinitesimal transfor-
mation. Hereafter, we will employ Einstein summation convention, i.e., aµbµ≡aµbµ≡/summationtextDim−1
µ=0ηµνaµbµwith vectors
aµandbµ(µ= 0,1,...,(Dim−1)), and the Minkowski metric: ηµν:= diag(−1,1,1,...,1/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright
Dim−1).
We define the variation operator of the action as follow:
δζI:=/integraldisplay
Ω′dDimx′L′(x′)−/integraldisplay
ΩdDimxL(x)
=/integraldisplay
dDimx′hΩ′(x′)L′(x′)−/integraldisplay
dDimxhΩ(x)L(x), (11)
where we characterize this variation by ζ, because this variation is derived from the infinitesimal transforma tions Eqs.
(9) and (10). Since the integration variable x′can be replaced by x, Eq. (11) is
δζI=/integraldisplay
dDimxhΩ′(x)L′(x)−/integraldisplay
dDimxhΩ(x)L(x)
=/integraldisplay
dDimx(hΩ′(x)−hΩ(x))L′(x)+/integraldisplay
dDimxhΩ(x)[L′(x)−L(x)]
=/integraldisplay
dDimxhδΩ(x)L′(x)+/integraldisplay
dDimxhΩ(x)[L′(x)−L(x)], (12)
whereδΩ := Ω′−Ω andhδΩ=−(∂µhΩ)δζxµ+O((δζx)2). Therefore, we obtain the following equation through partial
integration:
δζI=/integraldisplay
dDimxhΩ(x)/bracketleftbig
∂µ(L(x)δζxµ)+δL
ζL(x)/bracketrightbig
+O((δζx)2), (13)6
where we have introduced the so-called Lie derivative:
δL
ζφr(x) :=φ′
r(x)−φr(x)
=δζφr(x)−(∂µφr)δζxµ+O(δζx2), (14)
and we replacedL′byLdue toL′δζxµ=Lδζxµ+O(δζx2).
Hereafter, we assume that the action is invariant under the infinite simal transformations Eqs. (9) and (10). In the
case where the Lagrangian density is a function of φrand∂µφr, i.e.,L(x) =L[φr(x),∂µφr(x)], the Lie derivative of
the Lagrangian is given by
δL
ζL:=L′(x)−L(x)
=L[φ′
r(x),∂µφ′
r(x)]−L[φr(x),∂µφr(x)]
=/braceleftbigg∂L
∂φrδL
ζφr+∂L
∂(∂µφr)∂µδL
ζφr/bracerightbigg
=/parenleftbigg∂L
∂φr−∂µ∂L
∂(∂µφr)/parenrightbigg
δL
ζφr+∂µ/parenleftbigg∂L
∂(∂µφr)δL
ζφr/parenrightbigg
, (15)
and the variation of the action is calculated by
δζI=/integraldisplay
ΩdDimx/braceleftbigg/parenleftbigg∂L
∂φr−∂µ∂L
∂(∂µφr)/parenrightbigg
δL
ζφr+∂µ/parenleftbigg
Lδζxµ+∂L
∂(∂µφr)δL
ζφr/parenrightbigg/bracerightbigg
. (16)
If we require that δζxandδL
ζφrvanish on the surface ∂Ω, we obtain the Euler-Lagrange equation. On the other
hand, if we require that fields φrsatisfy the Euler-Lagrange equation, we obtain continuity equatio n∂µjµ= 0 with a
Noether current
jµ:=/parenleftbigg
Lδζxµ+∂L
∂(∂µφr)δL
ζφr/parenrightbigg
. (17)
Hereafter let us discuss an infinitesimal global U(1) ×SU(2) gauge transformation and infinitesimal translation and
rotation transformations, which are denoted by χin this paper. Variations in terms of χare defined by
δχxµ:= Γµ
νxν, (18)
δχφr:= iϑµν(ξµν)r′
rφr′ (19)
with an infinitesimal parameter ϑµν, and symmetry generators Γµ
νand (ξµν)r′
r.
1. Forthe globalU(1) ×SU(2) gaugetransformation, Γµ
ν≡0,ϑµν≡ϑµδµν, and (ξµν)r′
r≡δµν(ˆsµ)r′
r(µ,ν= 0,1,2,3;
r,r′= 1,2), where ˆs0:=/planckover2pi1/2, and ˆs1,2,3:=/planckover2pi1ˆσx,y,z/2 with the Planck constant h= 2π/planckover2pi1and Pauli matrices
ˆσx,y,z.
2. For the translation, Γµ
ν≡εµδµ
ν,ϑµν≡εµδµνand (ξµν)r′
r≡ˆpµδµνδr′
rwith an infinitesimal parameter εµand the
momentum operator ˆ pµ=−i/planckover2pi1∂µ(µ,ν= 1,2,3;r,r′= 1,2).
3. For the rotation, Γµ
ν≡ωµ
ν,ϑµν≡ωµν, and (ξµν)r′
r≡δr′
rxµˆpν, which corresponds to the angular momentum
tensor (µ,ν= 1,2,3;r,r′= 1,2).
For these transformations, equation ∂µδχxµ= 0 is satisfied. This can be seen explicitly as follows. The variations
of space coordinates of the global U(1) ×SU(2) and the translation transformations are given by δχxµ= 0 orδχxµ=
constant, respectively, and thus ∂µδχxµ= 0 is trivial. The variation of the rotation transformation is given by
δχxµ=ωµ
νxν, therefore∂µδχxµ=∂µωµ
νxν=ωµ
µ= 0.
We consider a variation of the Lagrangian density;
δζL:=L′(x′)−L(x)
=L′(x′)−L(x′)+L(x′)−L(x)
=δL
ζL(x′)+δζxµ∂µL+O((δζx)2). (20)
Note that Eq. (20) is correct for any infinitesimal transformation . Here we consider the global U(1) ×SU(2) gauge
transformation and/or the translation and rotation transforma tionsδχ. Because∂µδχxµ= 0, we obtain the following
equation:
δχL=δL
χL(x)+∂µ(L(x)δχxµ)+O((δx)2). (21)7
From Eqs.(13) and (21), one can see
δ(ζ=χ)I=/integraldisplay
dDimxδχL(x), (22)
where (ζ=χ) denotes that the type of the variation in Eq. (13) is restricted to the global U(1)×SU(2) or Poincare
transformations. (For simplicity we omitted the subscript Ω in the int egral). Finally, for ζ=χ, the variation of the
action is equal to the variation of Lagrangian. This fact will be used la ter in section V where the variation of the
Lagrangian density instead of the action will be considered.
B. Generalization of Noether’s theorem
Now, we wouldlike to introduce a Hopfalgebraforthe purposeofgen eralizingNoether’s theorem [8, 9, 13]. At first,
werewriteNoether’stheoreminsectionIIbyusingtheHopfalgebra ,andnext, weintroduceatwistedsymmetry [6,7].
For simplicity, we only consider the global U(1) ×SU(2) gauge symmetry and the Poincare symmetry. We assume tha t
the Lagrangian density is written as
L(x) =ψ†(x)ˆL(x)ψ(x) (23)
with a field ψ:=/parenleftbigg
ψ1
ψ2/parenrightbigg
, a Hermitian conjugate ψ†≡/parenleftbig
ψ1,ψ2/parenrightbig
, and an single-particle Lagrangian density operator
ˆL, which is a 2×2 matrix; the overline represents the complex conjugate. The act ion can be rewritten as
I=/integraldisplay
dDimx1dDimx2ψ†(x2)δ(x2−x1)ˆL(x1)ψ(x1)
= tr/integraldisplay
dDimx1dDimx2δ(x2−x1)ˆL(x1)ψ(x1)ψ†(x2)
= tr/integraldisplay
dDimx1/braceleftbigg/integraldisplay
dDimx2˜L(x1,x2)G(x2,x1)/bracerightbigg
= tr/integraldisplay
dDimx1/braceleftbigg
lim
x3→x1(˜L∗CG)(x1,x3)/bracerightbigg
, (24)
where “tr” represents the trace in the spin space, G(x1,x2) :=ψ(x1)ψ†(x2),˜L(x1,x2) :=δ(x1−x2)ˆL(x2), and∗C
represents the convolution integral:
(f∗Cg) :=/integraldisplay
dDimx3f(x1,x3)g(x3,x2) (25)
with smooth two-variable functions fandg.
The variation operator δχof the action can be also rewritten as
δχI= tr/integraldisplay
dDimx1dDimx2˜L(x2,x1)/bracketleftbig
iϑξψ(x1)ψ†(x2)−ψ(x1)ψ†(x2)iϑξ/bracketrightbig
= tr/integraldisplay
dDimx1dDimx2/bracketleftig
˜L(x2,x1)iϑξG(x1,x2)−iϑξ˜L(x2,x1)G(x1,x2)/bracketrightig
(26)
withϑξ≡ϑµν(ξµν); in addition, we assumed that the single-particle Lagrangian densit y operator is invariant under
the infinitesimal transformation δχ.
Here, we introduce Grassmann numbers θ1andθ2; an integral is defined by/integraltext
dθi(θj) =δij. The variation of the8
right-hand side of Eq.(26) can be rewritten as follow:
δχI=−itr/integraldisplay
dθ1dθ2dDimx1dDimx2/bracketleftig
θ1˜L(x2,x1)ϑθ2ξG(x1,x2)+ϑθ2ξθ1˜L(x2,x1)G(x1,x2)/bracketrightig
=−itr/integraldisplay
dθ1dθ2dDimx1dDimx2µ◦(µ⊗id)
◦/bracketleftig/parenleftig
θ1˜L(x2,x1)⊗ϑθ2ξ+ϑθ2ξ⊗θ1˜L(x2,x1)/parenrightig
⊗G(x1,x2)/bracketrightig
=−itr/integraldisplay
dθ1dθ2dDimx1dDimx2µ◦(µ⊗id)/bracketleftig
△(ϑθ2ξ)◦/parenleftig
θ1˜L(x2,x1)⊗G(x1,x2)/parenrightig/bracketrightig
≡ˆTr/bracketleftig
△(ϑθ2ξ)◦((θ1˜L)⊗G)/bracketrightig
, (27)
where⊗and◦represent a tensor product and a product of operators, respe ctively. The operator µdenotes the
transformation of the tensor product to the usual product µ:x⊗y/ma√sto→xy, and△represents a coproduct:
△(ζ) :=ζ⊗id+id⊗ζ, (28)
whereζand id represent a certain operator and the identity map, respect ively. These operators constitutes the Hopf
algebraaswillbeexplainedinthenextsection. Moreover,wehavede finedˆTr :=−itr/integraltext
dθ1dθ2dDimx1dDimx2µ◦(µ⊗id).
We emphasize here that the variation is written by the coproduct △, which is important to formulate the generalized
Noether theorem in the presence of the gauge potential. The copr oduct determines an operation rule of a variation
operator; for example, the coproduct (28) represents the Leib niz rule. A twisted symmetry transformation is given
by deformation of the coproduct.
We now sketch the concept of the twisted symmetry in deformation quantization [6, 7]. First, we assume that
the variation of action δζI0is zero, i.e., ζrepresents the symmetry transformation of the system corres ponding to
the actionI0. Next, we consider the action IAwith external gauge fields A. Usually, external gauge fields breaks
symmetries of I0, i.e.,δζIA/ne}ationslash= 0. Here we introduce a map: F(0/mapsto→A):I0/ma√sto→IA, which will be defined in section
IVF. The basic idea is to generalize the ”product” taking into accoun t the gauge interaction. Using this map, the
variation is rewritten as δζF(0/mapsto→A)I0/ne}ationslash= 0. On the other hand, when the twisted symmetry δt
ζ:=F(0/mapsto→A)δζF−1
(0/mapsto→A)can
be defined, we obtain the following equation:
δt
χIA=F(0/mapsto→A)δχF−1
(0/mapsto→A)IA
=F(0/mapsto→A)δχF−1
(0/mapsto→A)F(0/mapsto→A)I0
=F(0/mapsto→A)δχI0
= 0. (29)
Namely,δt
χcorresponds to a symmetry with external gauge fields. In the exp ression for the variation of action in
terms of the Hopf algebra Eq.(27), we can replace ∆ by ∆tcorresponding to the change from δχtoδt
χas shown in
section V. This is achieved by using the Hopf algebra and the deforma tion quantization, which will be explained in
sections III and IV, respectively. Therefore, we can generalize t he Noether’s theorem and derive the conservation law
even in the presence of the gauge field A.
III. HOPF ALGEBRA
Here we introduce a Hopf algebra. First, we rewrite the algebra usin g tensor and linear maps. Secondly, a coalgebra
is defined using diagrams corresponding to the algebra. Finally, we de fine a dual-algebra and Hopf algebra.
1. Algebra
We define the algebra as a k-vector space Vhaving product µand unitε. Here,krepresents a field such as the
complex number or real number. In this paper, we consider Vas the space of functions or operators. A space of
linear maps from a vector space V1to a vector space V2is written as Hom( V1,V2).
A productµis a bilinear map: µ∈Hom(V/circlemultiplytextV,V), i.e.,
µ:V/circlemultiplydisplay
V→V,(x,y)/ma√sto→xy, (30)9
and a unit is a linear map: ε∈Hom(k,V), i.e.,
ε:k→V, α/ma√sto→α·1 (31)
withx,y,xy∈Vandα∈k. Hereµandεsatisfies
µ((x+y)⊗z) =µ(x⊗z)+µ(y⊗z), µ(x⊗(y+z)) =µ(x⊗y)+µ(x⊗z), (32)
µ(αx⊗y) =αµ(x⊗y), µ(x⊗αy) =αµ(x⊗y), (33)
ε(α+β) =ε(α)+ε(β) (34)
withx,y,z∈Vandα,β∈k.
The product µhas the association property, which is written as µ◦(id⊗µ) =µ◦(µ⊗id). Because the left-hand
side and the right-hand side of the previous equation give the followin g equations:
µ◦(id⊗µ)(x⊗y⊗z) =µ(x⊗(yz)) =x(yz) (35)
and
µ◦(µ⊗id)(x⊗y⊗z) =µ((xy⊗z)) = (xy)z, (36)
for allx,y,z,xy,yz,xyz ∈V, thenµ◦(id⊗µ) =µ◦(µ⊗id) is equal to the association property x(yz) = (xy)z. This
property is illustrated as the following diagram:
V/circlemultiplytextV/circlemultiplytextV V/circlemultiplytextV
V/circlemultiplytextV V❄✲
✲❄µ⊗id
id⊗µ µ
µ/clockwise
Here/clockwisedenotes that this graph is the commutative diagram.
The unitεsatisfies the following equation: µ◦(ε⊗id) =µ◦(id⊗ε). Since the left-hand side and the right-hand
side of the previous equation give the following equations
µ◦(ε⊗id)(α⊗x) =µ⊗(α1V⊗x) =α1Vx=αx (37)
and
µ◦(id⊗ε)(x⊗α) =µ◦(x⊗α1V) =αx1V=αx (38)
for allx∈Vandα∈k, and∃1V∈V, then the unit can be written as µ◦(id⊗ε) =µ◦(ε⊗id). Note that
V∼k/circlemultiplytextV∼V/circlemultiplytextk, where∼represents the equivalence relation, i.e., a∼bdenotes that aandbare identified. This
property is illustrated as:
k/circlemultiplytextV V/circlemultiplytextk V/circlemultiplytextV
V✲
❄✛
❍❍❍❍❍❍❍ ❥✟✟✟✟✟✟✟ ✙ε⊗id id ⊗ε
µ
∼ ∼
Algebra is defined as a set ( V,µ,ε).10
2. Coalgebra
A coalgebra is defined by reversing the direction of the arrows in the diagrams corresponding to the algebra. Thus,
we will define a coproduct △∈Hom(V,V/circlemultiplytextV) and counit η∈Hom(V,V) with ak-vector space V.
A coproduct is a bilinear map from VtoV/circlemultiplytextV:
△:V→V/circlemultiplydisplay
V, (39)
and satisfies co-association property:
V/circlemultiplytextV/circlemultiplytextV V/circlemultiplytextV
V/circlemultiplytextV V✻✛
✛✻△⊗id
id⊗△ △
△/clockwise
Namely,
(id⊗△)◦△= (△⊗id)◦△ (40)
(Compare the diagram corresponding to the association property and that corresponding to the co-association prop-
erty).
A counitηis a linear map from Vto fieldk:
η:V→k, (41)
and satisfies the following diagram:
k/circlemultiplytextV V/circlemultiplytextk V/circlemultiplytextV
V✛
✻✲
❍❍❍❍❍❍❍ ❨
✟✟✟✟✟✟✟ ✯η⊗id id ⊗η
△∼ ∼
Namely,
(η⊗id)◦△= (id⊗η)◦△, (42)
whereV∼k/circlemultiplytextV∼V/circlemultiplytextk.
Since△andηare linear maps,△andηsatisfy
△(x+y) =△(x)+△(y),△(αx) =α△(x), (43)
η(x+y) =η(x)+η(y), η(αx) =αη(x) (44)
withx,y∈Vandα∈k. Note that V∼k/circlemultiplytextV∼V/circlemultiplytextkandV/circlemultiplytextV∼k/circlemultiplytextV/circlemultiplytextV∼V/circlemultiplytextk/circlemultiplytextV∼V/circlemultiplytextV/circlemultiplytextk.
A coalgebra is defined as a set ( V,△,η). For example, in the vector space D≡k/circleplustextk∂:={a0+a1∂|a0,a1∈k},
we define a coproduct △D(∂) =∂⊗1 + 1⊗∂and△D(1) = 1⊗1, and a counit ηD(∂) = 0 and ηD(1) = 1.
The set (D,△D,ηD) is coalgebra, because this set satisfies the equations: ( △D⊗id)◦△D= (id⊗△D)◦△Dand
(ηD⊗id)◦△D= (id⊗ηD)◦△D. Because the coproduct and counit are linear map, we only check th e above equations
with respect to x= 1 and∂.
Forx= 1,
(△D⊗id)◦△D(1) =△D(1)⊗1 = 1⊗1⊗1, (45)11
and
(id⊗△D)◦△D(1) = 1⊗△D(1) = 1⊗1⊗1. (46)
Therefore, (△D⊗id)◦△D(1) = (id⊗△D)◦△D(1). Moreover,
(ηD⊗id)◦△D(1) =ηD(1)⊗1 = 1⊗1, (47)
and
(id⊗ηD)◦△D(1) = 1⊗ηD(1) = 1⊗1. (48)
Therefore ( ηD⊗id)◦△D(1) = (id⊗ηD)◦△D(1).
Forx=∂,
(△D⊗id)◦△D(∂) =△D(∂)⊗1+△D(1)⊗∂=∂⊗1⊗1+1⊗∂⊗1+1⊗1⊗∂, (49)
and
(id⊗△D)◦△D(∂) =∂⊗△D(1)+1⊗△D(∂) =∂⊗1⊗1+1⊗∂⊗1+1⊗1⊗∂. (50)
Therefore, (id⊗△D)◦△D(∂) = (△D⊗id)◦△D(∂). Finally,
(ηD⊗id)◦△D(∂) =ηD(∂)⊗1+ηD(1)⊗∂= 1⊗∂=∂, (51)
and
(id⊗ηD)◦△D(∂) =∂⊗ηD(1)+1⊗ηD(∂) =∂⊗1 =∂. (52)
Therefore, ( ηD⊗id)◦△D(∂) = (id⊗ηD)◦△D(∂). Namely, the set ( D,△D,ηD) is the coalgebra. Note that △D(1)
corresponds to the product with a constant: a(fg) =a1(fg) =a(1f1g) =aµ◦△D(1)(f⊗g), where we have used the
coproduct△D(1) = 1⊗1atthefinalequalsign. Here f,andgaresmoothfunctions, 1isincludedinthefunctionspace,
anda∈k.△D(∂) represents the Leibniz rule: ∂(fg) = (∂f)g+f∂(g) =µ◦(∂⊗1+1⊗∂)◦(f⊗g) =µ◦△D(∂)(f⊗g),
where we have used the coproduct △D(∂) = 1⊗∂+∂⊗1 at the last equal sign. ηD(1) andηD(∂) represent the
filtering action to a constant function: 1 a=a=ηD(1)aand∂(a) = 0 =ηD(∂)a, respectively.
3. Dual-algebra and Hopf algebra
A dual-algebra is the set of an algebra and a coalgebra, i.e., the set of (V,µ,ε,△,η). On a dual-algebra, we define
a∗-product as
f∗g=µ◦(f⊗g)◦△ (53)
withf,g∈Hom(V,V). We define an antipode S∈Hom(V,V) which satisfies the following equation:
µ◦(id⊗S)◦△=µ◦(S⊗id)◦△=ε◦η, (54)
whereε◦ηcorresponds to the identity mapping, i.e., Sis an inverse of unit. For example, SDin the set
(D,µD,εD,△D,ηD) is defined as SD(1) = 1 and SD(∂) =−∂.
Forx= 1,
µD◦(id⊗SD)◦△D(1) =µD◦(1⊗1) = 1, (55)
and
µD◦(SD⊗id)◦△D(1) =µD◦(1⊗1) = 1. (56)
Therefore, we obtain µD◦(id⊗SD)◦△D(1) =µD◦(SD⊗id)◦△D(1) =εD◦ηD. For∂,
µD◦(id⊗SD)◦△D(∂) =µD◦(∂⊗1−1⊗∂) = 0, (57)12
and
µD◦(SD⊗id)◦△D(∂) =µD◦(−∂⊗1+1⊗∂) = 0. (58)
Therefore, we obtain µD◦(id⊗SD)◦△D(∂) =µD◦(SD⊗id)◦△D(∂) =εD◦ηD(∂). Namely, ( D,µD,εD,△D,ηD)
is the Hopf algebra.
A dual-algebra with an antipode S, i.e., (V,µ,ε,△,η,S), is called a Hopf algebra.
By using the approach similar to a coproduct and counit, we can defin e a codifferential operator Q∈Hom(V,V)
from a diagram of the differential ∂∈Hom(V,V). The differential ∂is the linear map:
∂:V→V, (59)
and satisfies Leibniz rule
∂◦µ=µ◦(id⊗∂+∂⊗id), (60)
which is illustrated as
V V
V/circlemultiplytextV V/circlemultiplytextV✻✛
✛✻∂
µ µ
(id⊗∂+∂⊗id)/clockwise
A codifferential operator Qis a linear map; Q:V→V, and satisfies the following diagrams:
V V
V/circlemultiplytextV V/circlemultiplytextV❄✲
✲❄Q
△ △
(id⊗Q+Q⊗id)/clockwise
Namely, a codifferential operator Qsatisfies△◦Q= (id⊗Q+Q⊗id)◦△. In section IVB2, the codifferential
operator will be introduced.
IV. DEFORMATION QUANTIZATION
In this section, we explain the deformation quantization using the no ncommutative product encoding the commu-
tation relationships. At first, in section IVA, we introduce the so-c alled Wigner representation and Wigner space,
and show that a product in the Wigner space is noncommutative. This product is called Moyal product and it
guarantees the commutation relationship of the coordinate and ca nonical momentum. Next, we add spin functions
and background gauge fields to the Wigner space, and rewrite the c oordinates of Wigner space as a set of spacetime
coordinates X, mechanical momenta p, and spins s:= (sx,sy,sz) (pincludes the background gauge fields). To gen-
eralize the Moyal product for the deformed Wigner space, which is a set of function defined on ( X,p,s), we explain
the general constructing method of the noncommutative produc t in section IVB; the noncommutative product is
the generalized Moyal product, which is called “star product”. This constructing method is given as a map from a
Poisson bracket in the Wigner space to the noncommutative produc t (see section IVB), and we see the condition of
this deformation quantization map in section IVB. This map is describe d by the path integral of a two-dimensional
field theory, which is called the topological string theory. In section IVC, we explain this topological string theory,
and in section IVD, we discuss the perturbative treatment of this t heory. In section IVE, we summarize the diagram
technique. Finally, in section IVF, we construct the star product in (X,p,s) space. We note that the star product
guarantees the background gauge structure.13
A. Wigner representation
We start with the introduction of the Wigner representation. From Equation (24), a natural product is the
convolution integral:
(f∗Cg)(x1,x2) :=/integraldisplay
dDimx3f(x1,x3)g(x3,x2), (61)
wheref,g∈Gwith a two spacetime arguments function space G. Here we introduce the center of mass coordinate X
and the relative coordinate ξas follows:
X≡(T,X) := ((t1+t2)/2,(x1+x2)/2), (62)
ξ≡(ξt,ξ) := (t1−t2,x1−x2). (63)
Moreover we employ the following Fourier transformation:
FT:f(x1,x2)/ma√sto→f(X,p) =/integraldisplay
dDimξe−ipµξµ//planckover2pi1f(X+ξ/2,X−ξ/2). (64)
Now we define the Wigner space: W:={FT[f]|f∈G}[14]. In this space, the convolution is transformed to the
so-called Moyal product [15, 16]:
(f ⋆Mg)(X,p) :=f(X,p)ei/planckover2pi1
2/parenleftBig← −∂X− →∂pν−← −∂p− →∂X/parenrightBig
g(X,p), (65)
because
F−1
T[f ⋆Mg] =/integraldisplaydDimp
(2π/planckover2pi1)Dimeipνξν//planckover2pi1/braceleftbigg
f(X,p)ei/planckover2pi1
2/parenleftBig← −∂Xν− →∂pν−← −∂pν− →∂Xν/parenrightBig
g(X,p)/bracerightbigg
=/integraldisplaydDimp
(2π/planckover2pi1)DimdDimξ1dDimξ2eipνξν//planckover2pi1/braceleftig
e−ipνξ1//planckover2pi1f(X+ξ1/2,X−ξ1/2)
×ei/planckover2pi1
2/parenleftBig← −∂Xν− →∂pν−← −∂pν− →∂Xν/parenrightBig
e−ipνξν
2//planckover2pi1g(X+ξ2/2,X−ξ2/2)/bracerightig
=/integraldisplaydDimp
(2π/planckover2pi1)DimdDimξ1dDimξ2eipν(ξν−ξν
1−ξν
2)//planckover2pi1
×f(X+ξ1/2,X−ξ1/2)e1
2/parenleftBig← −∂Xνξν
2−ξν
1− →∂Xν/parenrightBig
g(X+ξ2/2,X−ξ2/2)
=/integraldisplay
dDimξ1dDimξ2δ(ξ−ξ1−ξ2)
×f/parenleftbigg
X+ξ1+ξ2
2,X−ξ1−ξ2
2/parenrightbigg
g/parenleftbigg
X−ξ1−ξ2
2,X−ξ1+ξ2
2/parenrightbigg
=/integraldisplay
dDimx+dDimx−δ(ξ−x+)f/parenleftig
X+x+
2,x−/parenrightig
g/parenleftig
x−,X−x+
2/parenrightig
=/integraldisplay
dDimx−f(X+ξ/2,x−)g(x−,X−ξ/2)
=f∗Cg (66)
withξ1+ξ2≡x+andξ1−ξ2≡2(X−x−).
In the Wigner space, the position operator ˆ xµ=xµand the momentum operator ˆ pµ=−i/planckover2pi1∂µbecomesXµ⋆Mand
pµ⋆Mbecause
FT[ˆxµ
1g(x1,x2)] =/integraldisplay
dDimξe−i
/planckover2pi1pνξν(Xµ+ξµ/2)g(X+ξ/2,X−ξ/2)=/parenleftbigg
Xµ+i/planckover2pi1
2∂pµ/parenrightbigg
g(X,p) =Xµ⋆Mg(X,p),(67)
FT[ˆxµ
2g(x1,x2)] =/integraldisplay
dDimξe−i
/planckover2pi1pνξνg(X+ξ/2,X−ξ/2)(Xµ−ξµ/2) =g(X,p)⋆MXµ, (68)
FT[(ˆp1)µg(x1,x2)] =/integraldisplay
dDimξe−i
/planckover2pi1pνξν/planckover2pi1
i∂xµ
1g(x1,x2) =pµ⋆g(X,p), (69)14
and
FT[(ˆp2)µg(x1,x2)] =/integraldisplay
dDimξe−i
/planckover2pi1pνξν/planckover2pi1
i∂xµ
2g(x1,x2) =g(X,p)⋆pµ. (70)
The commutation relationship of operators is [ Xµ,pν]⋆M:=Xµ⋆Mpν−pν⋆MXµ= i/planckover2pi1δµ
ν, which corresponds to the
canonical commutation relationship of operators: [ˆ xµ,ˆpν] = i/planckover2pi1δµ
ν.
To add the spin arguments in W, we will employ the following bilinear map:
FM/ma√sto→FA=0:=ei/planckover2pi1
2(∂Xµ⊗∂pµ−∂pµ⊗∂Xµ)+i
2ǫabcsa∂sb⊗∂sc(71)
with∂saf:=faandf≡f0+/summationtext
a=x,y,zsafa. Note that the spin operator ˆs:= (ˆsx,ˆsy,ˆsz) is characterized by the
commutation relation [ˆ sa,ˆsb] = iǫabcˆsc(a,b,c=x,y,z) with the Levi-Civita tensor ǫabc, and the star product (71)
reproduces the relation, i.e., the operator ( sa⋆) satisfies [sa,sb]⋆= iǫabcsc.
To obtain the map F(0/mapsto→A):I0/ma√sto→IA, we introduce the variables transformation ( Xµ,pµ,s)/ma√sto→(Xµ,ˆpµ,s) where
ˆpµ=pµ−qAa
µ(Xν)sa+eAµ (72)
withq=|e|/mc2, the electric charge −e=−|e|, a U(1) gauge field Aµ, and a SU(2) gauge field Aa
µ. Their fields are
treated as real numbers, and the integral over pµcan be replaced by an integral over ˆ pµ. This transformation induces
the following transformations of differential operators:
∂Xµ⊗∂pµ−∂pµ⊗∂Xµ/ma√sto→∂Xµ⊗∂ˆpµ−∂ˆpµ⊗∂Xµ+q/parenleftig
∂XµˆAν−∂XνˆAµ/parenrightig
∂ˆpµ⊗∂ˆpν, (73)
ǫabcsa∂sb⊗∂sc/ma√sto→ǫabcsa∂sb⊗∂sc−qǫabcAb
µsa∂ˆpµ⊗∂sc−qǫabcAc
µsa∂sb⊗∂ˆpµ
+q2ǫabcsaAb
µAc
ν∂ˆpµ⊗∂ˆpν, (74)
whereˆAµ:=Aa
µsa−(e/q)Aµ.
We expandFA=0in terms of /planckover2pi1as
FA=0=∞/summationdisplay
n=0/parenleftbiggi/planckover2pi1
2/parenrightbiggn
Fn
A=0. (75)
We define the bilinear map FAcorresponding to the commutation relation in terms of the phase sp ace (Xµ,ˆpµ,s),
and expand it in terms of /planckover2pi1as
FA=∞/summationdisplay
n=0/parenleftbiggi/planckover2pi1
2/parenrightbiggn
Fn
A. (76)
From Eqs. (73) and (74), F1
Ais given as follows:
F1
A=∂Xµ⊗∂ˆpµ−∂ˆpµ⊗∂Xµ+qˆFµν∂ˆpµ⊗∂ˆpν+ǫabcsa∂sb⊗∂sc
−qǫabcsaAb
µ∂ˆpµ⊗∂sc+qǫabcsaAb
µ∂sc⊗∂ˆpµ (77)
withˆFµν:=∂XµˆAν−∂XνˆAµ+(q//planckover2pi1)εabcsaˆAb
µˆAb
ν. Note that µ◦F1
Ais the Poisson bracket.
A constitution method of higher order terms Fn
Awithn>1 is called a deformation quantization, which is given by
Kontsevich [17], as will be described in the next subsection.
B. Star product
In this subsection, we explain the Kontsevich’s deformation quantiz ation method [17]. We define a star product as
f ⋆g≡µ◦FA(f⊗g) =f·g+∞/summationdisplay
n=1νnβn(f⊗g) (78)
withν= i/planckover2pi1/2 [18, 19]. Here βn∈Hom(Vf⊗Vf,Vf) is called the two-cochain ( Vfrepresents the function space). We
require that the star product satisfies the association property (f ⋆g)⋆h=f ⋆(g⋆h), which limits forms of Fn≥1
A15
L∞L∞
d.g.L.aαβαβ
/g54
/g54 /g37/g372/g5422/g372/g40 /g40(a) (b) (c)
FIG. 2. Steps of the derivation of the deformation quantizat ion. (a): The image of the deformation quantization, which i s the
map from T2with the Jacobi identity to C2with the association property. (b): Enlargement of algebra s. The two vector space
T2and two cochain space C2generalize to multi-vector space Tand cochain space C, respectively. These spaces are compiled
in the d.g.L.a; finally, L∞algebra is introduced by using the d.g.L.a. (c): The deforma tion quantization is redefined as the
map on the L∞algebra.
andβn≥1. We note that the association property is necessary for the exist ence of the inverse with respect to the star
product. For example, the inverse of the Lagrangian is a Green fun ction, which always exists as ψψ†with a wave
functionψ.
Now, we define a p-cochain spaceCp:= Hom(V⊗p
f,Vf) withV⊗p
f≡Vf⊗Vf⊗···⊗Vf/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright
p, whereVfrepresents a
function space such as the Wigner space W; we define a multi-vector space Tk:= Γ(M,/logicalandtextkTM), whereMrepresents
a manifold such as a classical phase space (dimension d),TM:=/uniontext
p∈MTpMdenotes a tangent vector bundle with a
tangent vector space TpM≡{/summationtextd
iai(x)∂xi}atp∈M(xis a coordinate at p;airepresents a certain coefficient),/logicalandtextk
denotes ak-th completely antisymmetric tensor product, (for example, ∂i∧∂j=1
2!(∂i⊗∂j−∂j⊗∂i)∈/logicalandtext2TM), and
Γ represents the section; for example, Γ( M,TM) is defined as a set of tangent vector at each position p∈M. The
Poisson bracket{f,g}≡α(f⊗g) :=αij(x)(∂i∧∂j)(f⊗g) is element ofT1, whereαij=−αjiis called the Poisson
structure (i,j= 1,2,···,d).
The deformation quantization is the constitution method of higher o rder cochains βn≥2∈C2from the Poisson
bracketα∈T2. In other words, the deformation quantization is the following map F:
F:T2→C2
α/ma√sto→β≡/summationdisplay
n≥1νnβn, (79)
whereαsatisfies the Jacobi identity and βsatisfies the association property, as shown in Fig. 2(a).
In the following sections, we will generalize the two-cochain C2and the second order differential operator T2to the
so-calledL∞algebra(the definition is given in section IVB2). In the section IVB1, we will introduce the two-cochain
C2and second order differential operator T2, and thep-cochainCpandk-th order differential operator Tk. We will
show that these operators satisfy certain conditions, and CpandTkare embedded in a differential graded Lie algebra
(d.g.L.a) (the definition is shown in section IVB1). Moreover, in sectio n IVB2, the d.g.L.a will be embedded in the
L∞algebra (see Fig. 2(b)). In the L∞algebra, the Jacobi identity and the association property are com piled in the
following equation
Q(eγ) = 0, (80)
whereγ=αorβ, andQis called the codifferential operator, which will be introduced in sectio n IVB2. Namely,
in theL∞algebra, the deformation quantization is a map from α∈T2toβ∈C2holding the solution of Eq. (80)
(Figure 2(c)). Such a map is uniquely determined in the L∞algebra.
In this paper, we will identify the tensor product ⊗with the direct product ×, i.e.,V1/circlemultiplytextV2∼V1×V2:f⊗g∼(f,g)
withf∈V1andg∈V2(a∼bdenotes that aandbare identified; ( f,g) represents the ordered pair, i.e., it is a set of
fandg, and (a,b)/ne}ationslash= (b,a)).16
1. Cohomology equation
From Eq. (78), the association property is given by the following equ ation:
/summationdisplay
i+j=m
i,j≥0βi(βj(f,g),h)) =/summationdisplay
i+j=m
i,j≥0βi(f,βj(g,h)) (81)
withβ0(f,g)≡f·g. (The symbol “·” represents the usual commutative product, and βj∈C2, j= 0,1,···.) Because
β1is the Poisson bracket, which is bi-linear differential operator, we de fineβj(∈C2, j= 2,3,···) as a differential
operator on a manifold M; moreover, we also assume that p-cochains are differential operators and products of
functions.
Here,AandCk(A;A) represent a space of smooth functions on a manifold Mand a space of multilinear differential
maps fromA⊗ktoA, respectively. Degree of βk∈Ck(A;A) is defined by
deg(βk) :=kfork≥2. (82)
Now, we introduce a coboundary operator ∂C:Ck(A;A)→Ck+1(A;A) [20, 21];
(∂Cβk)(f0,···,fk/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright
k+1) :=f0βk(f1,···,fk/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright
k)+k/summationdisplay
r=1(−1)rβk(f0,···,fr−1·fr,···fk/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright
k)
+(−1)k−1βk(f0,···,fk−1/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright
k)fk (83)
withβk∈Ck(A;A); note that ∂2
C= 0, and thus, ∂Cis the boundary operator. The Gerstenhaber bracket is defined
as [,]C:Ck(A;A)⊗Ck′(A;A)→Ck+k′−1(A;A) [22]:
[βk,βk′]C(f0,f1,···,fk+k′−2/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright
k+k′−1)
:=k−1/summationdisplay
r=0(−1)r(k′−1)βk(f0,···,fr−1,βk′(fr,···,fr+k′−1),fr+k′,···,fk+k′−2/bracehtipupleft /bracehtipdownright/bracehtipdownleft /bracehtipupright
k)
−k′−1/summationdisplay
r=0(−1)(k−1)(r+k′−1)βk′(f0,···,fr−1,βk(fr,···,fr+k−1),fr+k,···,fk+k′−2/bracehtipupleft /bracehtipdownright/bracehtipdownleft /bracehtipupright
k′),
(84)
whereβk∈Ck(A;A) andβk′∈Ck′(A;A). Note that ∂2
C= 0, and thus, ∂Cis the boundary operator.
By using the coboundary operator and the Gerstenhaber bracke t, Eq. (81) is rewritten as
∂Cβm=−1
2/summationdisplay
i+j=m
i,j≥0[βi,βj]C (85)
withβj∈C2(A;A) (j= 1,2,···). For example, Eq. (81) for m= 0,1,2 is given as:
(f·g)·h=f·(h·g) for m= 0,
{f·g,h}+{f,g}·h={f,g·h}+f·{g,h} form= 1,
β2(f·g,h)+{{f,g},h}+β2(f,g)·h=β2(f,g·h)+{f,{g,h}}+f·β2(g,h) form= 2,(86)
where we have used β0(f,g) :=f·gandβ1(f,g)≡{f,g}. The coboundary operator for β∈C2(A;A) is given by:
(∂Cβ)(f,g,h) =f·β(g,h)−β(f·g,h)+β(f,g·h)−β(f,g)·h; (87)
moreover, the Gerstenhaber bracket in terms of βi,βj∈C2(A;A) is given by
[βi,βj]C(f,g,h) =βi(βj(f,g),h)−βi(f,βj(g,h))+βj(βi(f,g),h)−βj(f,βi(g,h)). (88)
Using the above Eqs. (86-88), we can check the equivalence betwe en Eq. (81) and Eq. (85).17
Equation (85) is called the cohomology equation, and the star produ ct is constructed by using solutions of the
cohomology equation. If we add Eq. (85) with respect to m= 0,1,2,···, we obtain the following equation:
∂Cβ+1
2[β,β]C= 0 (89)
withβ≡/summationtext∞
j=0βj;β,βj∈C2(A;A),j= 0,1,2,....
Here, we identify the vector fields ∂i,∂j∈TMwith anti-commuting numbers ˜ ηi,˜ηj(˜ηi˜ηj=−˜ηj˜ηi),i,j= 1,2,...,d;
thus the Poisson bracket αij(∂i∧∂j)/2 is rewritten by α=αij˜ηi˜ηj/2. Now, we define the Batalin-Vilkovisky (BV)
bracket:
[α1,α2]BV:=−d/summationdisplay
i=1/parenleftigg
α1← −∂
∂xi− →∂α2
∂˜ηi−α1← −∂
∂˜ηi− →∂α2
∂xi/parenrightigg
(90)
withα1,α2∈T2. By using the BV bracket, the Jacobi identity is rewritten as
∂BVα+1
2[α,α]BV= 0, (91)
withα1,α2∈T2; forα=αij˜ηi˜ηj,α← −∂/∂xl=− →∂α/∂xl:= (∂xlαij)˜ηi˜ηjand− →∂α/∂˜ηl=−α← −∂/∂˜ηl:=αij(δil˜ηj−˜ηiδjl).
By using the BV bracket, the Jacobi identity is rewritten as
∂BVα+1
2[α,α]BV= 0, (92)
where∂BV≡0, i.e.,∂2
BV= 0.
Now, we generalize the differential ∂BVand BV-bracket [ ,]BVforαk∈Tkandαk′∈Tk′as follows (Tk≡
Γ(M,/logicalandtextkTM)):
∂BV:Tk→Tk+1
∂BV:= 0, (93)
[,]BV:Tk/circlemultiplydisplay
Tk′→Tk+k′−1
[αk,αk′]BV:=−d/summationdisplay
i=1/parenleftigg
αk← −∂
∂˜ηi− →∂αk′
∂xi−αk← −∂
∂xi− →∂αk′
∂˜ηi/parenrightigg
(94)
withαk= (αk)i1,···,ik(x)ηi1∧···∧ηik∼(αk)i1,···,ik(x)˜ηi1···˜ηik, andαk′= (αk′)i0,···,ik′(x)ηi0∧···∧ηik′∼
(αk′)i0,···,ik′(x)˜ηi0···˜ηik′; degree ofα∈Tkis defined by
deg(α) =k−1, α∈Tk. (95)
The cochain algebra is defined by the set of the differential operato r∂C, the Gerstenhaber bracket [ ,]Cand
C:=/circleplustext∞
k=2Ck, i.e., (∂C,[,]C,C); in addition, the multi-vector algebra is defined by the set of the diff erential operator
∂BV:= 0, BV bracket [ ,]BVandT:=/circleplustext∞
k=1Tk, i.e., (∂BV,[,]BV,T). The cochain algebra and the multi-vector
algebra satisfy the following common relations:
∂2= 0, (96)
∂[γ1,γ2] = [∂γ1,γ2]+(−1)deg(γ1)[γ1,∂γ2], (97)
[γ1,γ2] =−(−1)deg(γ1)deg(γ2)[γ2,γ1] (98)
[γ1,[γ2,γ3]]+(−1)deg(γ3)(deg(γ1)+deg(γ2))[γ3,[γ1,γ2]]+(−1)deg(γ1)(deg(γ2)+deg(γ3))[γ2,[γ3,γ1]] = 0 (99)
withγ1,γ2,γ3∈G≡(CorT),∂≡∂(CorBV), and [,]≡[,](CorBV). Therefore, the two algebra can be compiled
in the so-called the differential graded Lie algebra (d.g.L.a) ( ∂,[,],G), whereG:=/circleplustext∞
k=1Gkis a graded k-vector18
space withGkhas a degree deg( x)∈Z(x∈Gk;Zis the set of integers), and d.g.L.a. has the linear operator ∂and
the bi-linear operator [ ,]:
∂:Gk→Gl, xk∈Gk, xl∈Gl,
deg(∂xk) = deg(xk)+1 = deg( xl), (100)
[,] :Gk/circlemultiplydisplay
Gl→Gm, xk∈Gk, xl∈Gl, xm∈Gm,
deg([xk,xl]) = deg(xk)+deg(xl) = deg(xm), (101)
where∂and [,] satisfy Eqs. (96), (97), (98) and (99). In d.g.L.a., Eqs. (89) and (92) are compiled in the so-called
Maurer-Cartan equation [23]:
∂γ+1
2[γ,γ] = 0 (102)
withγ∈G. Therefore, the deformation quantization Fis a map:
F:G→G, γ1/ma√sto→γ2,
∂γi+1
2[γi,γi] = 0, i= 1,2. (103)
Namely, the deformation quantization is a map holding a solution of the Maurer-Cartan equation (102). In the
section IVB2, we will introduce a L∞algebra, and will redefine the deformation quantization; in the L∞algebra,
the Maurer-Cartan equation (102) is rewritten as Q(eγ) = 0 (Qandeγwill be defined in IVB2).
2.L∞algebra
Now we define a commutative graded coalgebra C(V).
First, wedefineaset( V,△,τ,Q), whereV:=/circleplustext
n=1,2,···V⊗nwithagraded k-vectorspace V⊗n(n= 1,2,...),△and
Qrepresent the coproduct and codifferential operator, respect ively; moreover, τdenotes cocommutation (definition
is given later). The coproduct, cocommutation and codifferential o perator satisfy the following equations:
(△⊗id)◦△= (id⊗△)◦△, (104)
τ△=△, (105)
△◦Q= (id⊗Q+Q⊗id)◦△, (106)
τ(x⊗y) := (−1)degco(x)degco(y)y⊗x, (107)
with degco(x) := deg(x)−1, wherex∈V⊗deg(x)andy∈V⊗deg(y).Qrepresents a codifferential operator adding one
degree:Q∈Hom(V⊗m,V⊗(m+1)) with degco(Q(x)) = degco(x)+1 forx∈V⊗m,∃m∈Z+(the explicit form of Qis
given later; Z+:={i|i>0, i∈Z}).
By usingτ, we define the commutative graded coalgebra C(V) from (V,△,τ,Q); the identify relation ∼is defined as
x⊗y∼(−1)degco(x)degco(y)y⊗x, i.e.,x⊗yand (−1)degco(x)degco(y)y⊗xare identified. Now, we define the commutative
graded tensor algebra:
C(V) :=V/∼ ≡{[x]|x∈V}, (108)
where [x] ={y|y∈V, x∼y}, and degco(x1⊗x2⊗···⊗xn) = degco(x1) + degco(x2) +···+ degco(xn) with
x1⊗x2⊗···⊗xn∈V⊗degco(x1)⊗V⊗degco(x2)⊗···⊗V⊗degco(xn); a product inC(V) is defined by xy:= [x⊗y].
Namely, inC(V),
x1x2···xixi+1···xn= (−1)degco(xi)degco(xi+1)x1x2···xi+1xi···xn (109)
withn≥2. (Let us recall that the derivation of the exterior algebra from t he tensor space;VandC(V) correspond
to the tensor space and the exterior algebra, respectively.)
Moreover, in the case that Q2= 0, the commutative graded coalgebra C(V) is called the L∞algebra. For the L∞
algebra, the coproduct and codifferential operator are uniquely d etermined by using multilinear operators:
lk: (V⊗k∈C(V)))→V∈C(V) (110)
degco(lk(x1···xk)) = degco(x1)+···+degco(xk)+1 (111)19
as follows:
△(x1···xn) =/summationdisplay
σn−1/summationdisplay
k=1ε(σ)
k!(n−k)!(xσ(1)···xσ(k))⊗(xσ(k+1)···xσ(n)), (112)
Q=∞/summationdisplay
k=1Qk, (113)
Qk(x1···xn) =/summationdisplay
σε(σ)
k!(n−k)!lk(xσ(1)···xσ(k))⊗xσ(k+1)⊗···⊗xσ(n),
(114)
whereε(σ) represents a sign with a replacement σ:x1x2···xn/ma√sto→xσ(1)xσ(2)···xσ(n). From the condition Q2= 0,
we can identify ( l1, l2) with (∂,[,]) in d.g.L.a. If we put l3=l4=···= 0,Q(eα) = 0 forα∈Vis equal to the
Maurer-Cartan equation Eq. (102) in d.g.L.a [17], where
eα≡1+α+1
2!α⊗α+··· (115)
withα⊗n⊗1≡1⊗α⊗n≡α⊗nforn= 1,2,···. Therefore, the deformation quantization is a map:
F:C(V)→C(V),
γ1/ma√sto→γ2 (116)
with
Q(eγi) = 0, i= 1,2. (117)
To constitute such a map F, we introduce the L∞mapF, which is defined as the following map holding degrees
of coalgebra:
F:C(V)→C(V), v1,v2∈C(V),
v1/ma√sto→v2,
degco(v1) = degco(v2); (118)
moreover, the L∞map satisfies the following equations:
△◦F= (F⊗F)◦△, (119)
Q◦F=F◦Q. (120)
A form of such a map is limited as [17]:
F=F1+1
2!F2+1
3!F3+···, (121)
Fl:C(V)→V⊗l(⊂C(V))
Fl(x1···xn) =/summationdisplay
σ/summationdisplay
n1,···,nl≥1
n1+···+nl=nε(σ)
n1!···nl!
·Fn1(xσ(1)···xσ(n1))⊗···⊗ Fnl(xσ(n−nl+1)···xσ(n)), (122)
where Fnis a map fromC(V) toV(⊂C(V)) holding degrees;
Fn: V⊗n(⊂C(V))→V(⊂C(V))
x1⊗···⊗xn/ma√sto→x′,
degco(x1)+···+degco(xn) = degco(x′). (123)
Here we define β:=/summationtext∞
n=11
n!Fn(α···α), which satisfies F(eα) =eβ. The map Fholds solutions ofMaurer-Cartan
equationsQ(eα) = 0 andQ/parenleftbig
eβ/parenrightbig
= 0; from
Q(eβ)≡Q◦F(eα) (124)20
and the definition of the L∞map:Q◦F=F◦Q, we obtain the following equation:
Q(eβ) =F◦Q(eα) = 0, (125)
which means that the L∞map transfers a solution of the Maurer-Cartan equation from ano ther solution.
Now, we return to the deformation quantization. The multi-vector spaceT, is embedded in C(V);C(V) =
(T,△T,τ,QT), where△T(x1x2) :=x1∧x2forx1,x2∈T, (QT)1:=∂BV≡0, (QT)2:= [,]BV, and (QT)l:= 0 for
l= 3,4,...;τreplaces the wedge product “ ∧” with the product “ ·”. For the cochain space C, it is also embedded in
C(V);C(V) = (C,△C,τ,QC), where△C(x1x2) :=x1∧x2forx1,x2∈C, (QC)1:=∂C, (QC)2:= [,]C, and (QC)l:= 0
forl= 3,4,....
The star product is given by f ⋆g=f·g+β(f⊗g), which is identified as the map F0+F1withF0:=µ◦.
Here we summarize the main results of the succeeding sections witho ut explaining their derivations. The map
F1is given by a path integral of a topological field theory having super fi elds:X:= (X1,...,XN); and scalar
fields:ψ:= (ψ1,...,ψN),λ:= (λ1,...,λN), andγ:= (γ1,...,γN); and one-form fields: θ:= (θ1,θ2,...,θm),
A:= (A1,...,A N),A+:= (A+1,...,A+N), andη:= (η1,...,ηN); and Grassmann fields ci:= (c1,...,cN); on a disk
Σ ={z|z=u+iv, u,v∈R, v≥0}[17, 24, 25]. These fields are defined in section IVC. Using these field s, the
mapFn:V⊗n→Vis given as follows:
Fn(α1,···,αn)(f1⊗···⊗fm)(x) =/integraldisplay
ei
/planckover2pi1S0
ghi
/planckover2pi1Sα1···i
/planckover2pi1SαnOx(f1,...,f m) (126)
for any function f1,...,f m, which depend on x; in this paper, xrepresents the coordinate in the classical phase space.
Hereα1,α2,...,α n∈V, andmis defined by degco(αi)+2, which is common and independent of i(i= 1,2,...,n).
The operatorOxis defined as
Ox(f1,...,f m) :=/integraldisplay
[f1(X(t1,θ1))···fm(X(tm,θm))]δx(ψ(∞)) (127)
≡/integraldisplay
1=t1>t2>···>tm=0f1(ψ(t1))m−1/productdisplay
k=2∂ik/bracketleftbig
f(ψ(tk))A+ik(tk)/bracketrightbig
fm(ψ(0))δx(ψ(∞)) (128)
form,δx(ψ(t)) :=/producttextd
i=1δ(ψi−xi)γi(t), andt∈∂Σ, where
S0
gh:=/integraldisplay
Σ/bracketleftbig
Ai∧dψi−∗Hdγi∧dci−λid∗HAi/bracketrightbig
(129)
with a Hodge operator ∗H:∧k→∧2−k, (k= 0,1,2); we will introduce the explicit definition in section IVD2.
Moreover, for αr:=αi1,···,inrr(X)∂i1∧···∧∂inr(nr>1 is an integer number; degco=nr−2),
Sαr:=/parenleftbigg/integraldisplay
Σ/integraldisplay
d2θ1
nrαi1···inrr(X)ηi1···ηinr/parenrightbigg/vextendsingle/vextendsingle/vextendsingle/vextendsingle
Φ∗=∂ϕ, (130)
where the subscript Φ∗=∂ϕmeans that the fields ( X,η,A+) go to (ψ,A,0). These results lead to the diagram
technique in section IVE and the explicit expression of the star prod uct in section IVF.
C. Topological string theory
In this section, we expound the fields: A,ψ,c,γ,λ,θ,η,A+, andX. The simplest topological string theory is
defined the following action:
S0:=/integraldisplay
Σd2σǫµνAµ,i∂νψi=−1
2/integraldisplay
Σd2σǫµνFµν,iψi(131)
with local coordinates σ= (σ1,σ2) on a disk Σ (we consider that the disk is the upper-half plain in the com plex
one, i.e., Σ :={z|z=u+iw;w≥0;u,w∈R}), whereAµ,i(σ) andψi(σ) are U(1) gauge fields and scalar fields,
respectively; Fµν,i(σ) is a gauge strength ( µ= 1,2 andi= 1,...,N). The other fields c,γ,λ,θ,η, andA+are
introduced in section IVC1; we discuss the gauge fixing method using the so-called BV-BRST formalism [26, 27]
(where the BV refers to Batalin and Vilkovisky; BEST refers to Becc hi, Rouet, Stora and Tyutin). In section IVC2,
we discuss the gauge invariance of the path integral, and introduce the SD operator. In section IVC3, we see that
correspondence of the deformation quantization and topological string theory.21
1. Ghost fields and anti-fields
Here, we quantize the action (131) using the path integral. Roughly speaking, the path integralis the Gaussintegral
around a solution of an equation of motion. In many cases, a genera l actionShas no inverse. Therefore, we will add
some extra fields, and obtain the action Sghhaving inverse, which is called as the quantized action.
Now, we discuss a general field theory. We assume that a general a ctionSis a function of fields φi, i.e.,S=S[φi];
each fieldφiis labeled by a certain integer number, which is called as a ghost number gh(φi) (it is defined below).
φi
Cdenotes that the fields fixed on the solution of the classical kinetic e quation:δS0/δφi= 0, and the subscript of
the fields represents a number of fields. Because the Gauss integr al is an inverse of a Hessian, a rank of the Hessian
should be equal to the number of the fields. Here a Hessian is defined by:
K[φi,φj] :=− →δ
δφiS← −δ
δφj, (132)
where− →δ
δφiφj1φj2···φin:=δj1
iφj2···φj2+(−1)j1iφj1δj2
i···φjn
i+···+(−1)i(j1+j2+···+j(n−1))φj1φj2···φj(n−1)δjn
i, and
φi1φi2···φin← −δ
δφj:=φi1φi2···φi(n−1)δin
j+(−1)injφi1φi2···δi(n−1)
jφin+(−1)(i2+···+in)jδi1
jφi2···φin; for a boson φi,i
is a ghost number gh( φi); for a fermion φi,iis gh(φi)+1.
We define the rank of the Hessian Kand the number of the fields φiby♯Kand♯φi, respectively. Generally
speaking,♯K <♯φi, because an action has some symmetries δRφi:=Ri
jφjwith nontrivial symmetry generators Ri
j,
where is satisfies the following equation:
S← −δ
δφiRi
j= 0 (133)
withRi
j|φk=φkc/ne}ationslash= 0. The nontrivial symmetry generator decrease the rank of Hes sian from the number of fields. To
define the path integral, we should add ( ♯φi−♯K) virtual fields [26–28]. The additional fields are called as ghost fields
Φα1and antifields Φ∗
αl(l= 0,1), and these fields are labeled by ghost numbers. For Φα1, the ghost number is defined
by gh(Φα1) := 1. The fields and ghost fields have antifields Φ∗
αl. The antifields corresponding to φ≡Φα0and Φα1
are described as Φ∗
α0and Φ∗
α1, respectively. The ghost number of Φ∗
αlis defined by gh(Φ∗
αl) =−l−1. Statistics of
the anti-fields is opposite of fields, i.e., if the fields are fermions(boso ns), the anti-fields are bosons(fermions). (Here
we only consider the so-called irreducible theory. For a general the ory, see references [26–28].)
Using these fields, we will transform the action S[Φα0]/ma√sto→Sgh[Ψ], where Ψ := (Φαl,Φ∗
αl) withl= 0,1 and
αl∈Z+:={i|i>0, i∈Z}(Zrepresents the set of integers), Φα0:=φiare fields, Φα1represents ghost fields, and
Φ∗
αldenotes anti-fields of the fields Φαl. Hereafter we write a function space created by the fields and ant i-fields as
C(Ψ). It is known that Sghis given by
Sgh=S+Φ∗
α0Rα0α1Φα1+O(Ψ3). (134)
Note that the anti-fields will be fixed, and ♯Ψ =/summationtext
l♯Φαl(see section IVC3).
For the topological string theory, the fields φαare U(1) gauge fields Ai,µand scalar fields ψiwithi= 1,...,Nand
µ= 1,2; namely, Φα0≡φα:= (Ai,µ,ψi). Since♯Ai,µ= 2Nand♯ψi=N, the fields number ♯φαis 3N. The action
(131) has the U(1) gauge invariance:
δ0Aµ,i=∂µδj
iχj, (135)
δ0ψi= 0, (136)
δ0χi= 0 (137)
withχirepresentsa scalarfunction ( i= 1,...,N). Therefore, the topologicalstring theory has 2 Nlinear-independent
nontrivial symmetry generators. Here we replace the scalar fields
chiiwith ghost fields ci(BRST transformation). Moreover, we add antifields A∗
i,µ; since the gauge transformation
does not connect to ψand the other fields, we does not add ψ∗(the space of fields and ghost fields has 2 Nsymmetry
generators, and the space of anti-fields and the anti-ghosts also have 2Nsymmetry generators corresponding to U(1)
gauge symmetry; see Figure 3):
R(µ,i)
β=∂µδi
β,(β= 1,2,...,N). (138)22
FIG. 3. The Hessian matrix: K[Ψα(σ1),Ψβ(σ2)] :=δ
δΨα(σ2)Sδ
δΨβ(σ1)The first column and first raw represent the right-hand
side and the left-hand side of variation functions, respect ively.ˆ∂σj:=δ(σ−σ1)∂σjδ(σ−σ2) represents a non-trivia Noether
current ( j= 1,2). The Hessian is block diagonal matrix; the ranks of the upp er left and the lower right parts are 2 = 3 −1.
Therefore, the total rank of the Hessian is 2+2 = 4 ( iis fixed).
In this case, ♯K(φα,φβ) = 3N−2N; on the other hand, the action is a function of 3 Nfields (Aµ,i,ψi),Nghost fields
ciand 2Nanti-fieldsA∗
i,µ. Therefore, a rank of the Hessian corresponding to ( S0)ghis calculated by
rankK(Ψ,Ψ)|Ψc= rankK(φ,φ)|Ψc+♯ci+♯A∗
i,µ
=N+N+2N
= 4N. (139)
Since♯Φ = 4N(antifields will be fixed), the field number of the path-integral of ( S0)ghis equal to the rank of the
Hessian of ( S0)gh; hence, the path-integral of the action ( S0)ghbecome well-defined.
Finally, the gauge invariance action is written by
(S0)gh=S0+/integraldisplay
Σ(Ai)+∧δ0Ai (140)
withAi:=Aµ,idσµandδ0Aµ,i=∂µδj
icj, where we define Φ+
αusing a Hodge operator ∗H: Φ+
α≡ ∗HΦ∗
α(the
definition of the Hodge operator is depend on the geometry of the d isk Σ; we will introduce the explicit definition in
section IVD2), which is also called as the anti-field.
2. Condition of gauge invariance of classical action
In this section, we will add interaction terms: Sgh:= (S0)gh+g(S1)gh+···, wheregrepresents an expansion
parameter, and we will see that Sghis uniquely fixed except a certain two form αby a gauge invariance condition.
Note thatαsatisfies the Jacobi’s identity. Therefore, we can identify αwith the Poisson bracket.
First we discuss the gauge invariant condition. If we identify the field s and anti-fields with coordinates qand
canonical momentum p, i.e., (Φαi,Φ∗
αi)↔(qαi,pαi), and we also identify the action Sand the Hamiltonian H:
S↔H. In the analytical mechanics, δam:={H,}represents a transform along the surface H(q,p) = constant,
i.e.,δamholds the Hamiltonian. Similarly, we can define a gauge transformation , which holds the action S, using the
Poisson bracket in the two-dimension field theory. It is known as the Batalin-Vilkovisky (BV) bracket [26, 27]; the
definitions of the bracket are
{f,g}BV:=/summationdisplay
αi
i=0,1,.../parenleftbiggδf
δΦαiδg
δΦ∗αi−δf
δΦ∗αiδg
δΦαi/parenrightbigg
(141)
withf, g∈C(Ψ).
The BV bracket has the ghost number 1, then a BV-BRST operator δBV:={S,}BVadds one ghost number. The
BV bracket satisfies the following equations:
{f,g}BV=−(−1)(gh(f)+1)(gh( g)+1){g,f}BV, (142)
(−1)(gh(f)+1)(gh( h)+1){f,{g,h}BV}BV+cyclic = 0 , (143)
{f,gh}BV={f,g}BVh+(−1)(gh(f)−1)gh(g)g{f,h}BV (144)23
withf, g, h∈C(Ψ).
Using the BV-BRST operator, the gauge invariance of action Sis written as δBVS= 0, i.e.,
{S,S}BV= 0, (145)
which is called the classical master equation. We use this equation and Eqs. (142) and (143); we obtain δ2
BV= 0,
which corresponds to the condition of the BRST operator: δ2
BRST= 0 (δBRSTis the BRST operator). Therefore, the
DV-BRST operator is the generalized BRST one.
Next, we discuss generalization of the topological field theory. Let us write a generalized action Sghas
Sgh= (S0)gh+g(S1)gh+g2(S2)gh+···, (146)
wheregis an expansion parameter. Using gauge invariance condition (145), (Sn)gh(n= 1,2,···) is given by a
solution of the following equation:
∂n
∂gn{Sgh,Sgh}BV/vextendsingle/vextendsingle/vextendsingle/vextendsingle
g=0= 0. (147)
The general solution is given by [25]
(S1)gh=/integraldisplay
Σd2σ/bracketleftigg
1
2αij(AiAj−2ψ+
icj)+∂αij
∂ψk/parenleftbigg1
2(c+)kcicj−(A+)kAicj/parenrightbigg
+1
4∂2αij
∂ψk∂ψl(A+)k(A+)lcicj/bracketrightigg
, (148)
(Sn>1)gh= 0 (149)
with (A+)i≡ ∗HA∗
µ,i=dσµεµνA∗
ν,i,ψ+
i≡∗Hψ∗
i=εµνdσµ∧dσνψ∗
iand (c+)i≡∗H(c∗)i=εµνdσµ∧dσν(c∗)i
(εµν=−ενµ, ε12= 1), where αijis a function of ψ, and satisfies the following equation:
∂αij
∂ψmαmk+∂αjk
∂ψmαmi+∂αkl
∂ψmαmj= 0. (150)
Here, if we identify ψiwithxi, this equation is the Jacobi identity of Poisson bracket. Therefor e, we can identify the
Poisson bracket with the topological string theory.
3. Gauge invariance in path integral
Now we discuss the path integral of the topological string theory/integraltext
DΦV(Ψ) withV(Ψ) =Oei
/planckover2pi1S, and an observable
quantity operator O. Note that this path-integral does not include integrals in terms of the anti-fields. Therefore,
we must fix the anti-fields; then, we consider that the anti-field Φ∗is a function of the field Φ, i.e., Φ∗= Ω(Φ) and
Ω∈C(Φ = Ψ) Namely, the path integral is defined by
/integraldisplay
DΦV(Ψ)/vextendsingle/vextendsingle/vextendsingle/vextendsingle
Φ∗=Ω. (151)
A choice of Ω(Φ) is corresponding to the gauge fixing in the gauge the ory. The path integral must be independent
to the gauge choice (gauge invariance). To obtain a gauge invariant condition, we take the variation of the path
integral in terms of anti-fields, and obtain the following gauge invaria nt condition [29]:
△SDV(Ψ) = 0, (152)
where we have introduced the Schwinger-Dyson (SD) operator:
△SD:=/summationdisplay
αl(−1)αlδ
δΦαlδ
δΦ∗αl, (153)24
where (−1)αlis defined as follows: if Φαlrepresents a boson, ( −1)αl= (−1)gh(Φαl); if Φαlrepresents a fermion,
(−1)αl= (−1)(gh(Φαl)+1). Equation (152) is called the quantum master equation. It is known t hat the following two
conditions are equivalence:
△SDV(Ψ) = 0⇐⇒Ω =− →δϕ
δΦa,∃ϕ, (154)
whereϕis called the gauge-fixing fermion (an example will be shown later).
To performthe path integral, we generalizethe classicalaction Sghtoa quantum action W=Sgh+i/planckover2pi1W1+(i/planckover2pi1)2W2+
···. The correction terms Wn(n= 1,2,...) are calculated from the master equation:
△SDei
/planckover2pi1W= 0, (155)
or
{Sgh,Sgh}BV= 0, (156)
{W1,Sgh}BV+i/planckover2pi1△SDSgh= 0, (157)
{W2,Sgh}BV+i/planckover2pi1△SDW1+1
2{W1,W1}BV= 0, (158)
···
In the case where △SDSgh= 0, we can put W1=W2=···= 0. Fortunately, the topological string theory satisfies
△SDSgh= 0. Therefore, we do not have to be concerned about the quantu m correction of the action.
Finally, we consider the gauge fixing. Here we employ the Lorentz gau ge:
d∗HAi= 0, (159)
and we add the integral of the Lorentz gauge to Sgh. However, the path integral should hold gauge invariance, i.e., the
path integral should be independent of gauge fixing term. Then, th e gauge fixing can be written gauge-fixed fermion:
ϕ:=/integraldisplay
Σγi(d∗HAi) =−/integraldisplay
Σdγi∗HAi, (160)
where we introduced Nfieldsγi(i= 1,2,...,N), and anti-fields γ+
iare given by
γ+
i=− →∂ϕ
∂γi=d∗HAi. (161)
Now, we employ the Lagrange multiplier method, and introduce Nscalar fields λi. The gauge-fixed action is written
by
Sgf=Sgh−/integraldisplay
Σγid∗HAi (162)
=Sgh−/integraldisplay
Σλiγ+
i. (163)
The other anti-fields are also fixed by this gauge-fixing fermion:
ψ+
i=c+
i=λ+
i= 0, (164)
A+
i=∗Hdγi. (165)
Gauge fixed action Sgfis written by
Sgf=/integraldisplay
Σ/bracketleftigg
Ai∧dψi+1
2αijAi∧Aj−∗Hdγi∧/parenleftbigg
dci+∂αkl
∂ψiAkcl/parenrightbigg
−1
4∗Hdγi∧∗Hdγj∂2αkl
∂ψi∂ψjckcl−λid∗HAi/bracketrightigg
. (166)25
Here we perform the following variable transformations:
Xi:=ψi+θµA∗
µ−1
2θµθνc+i
µν, (167)
ηi:=ci+θµAi,µ+1
2θµθνψ+
i,µν, (168)
whereθµϑν=−θνθµ; gh(θµ) = 1. For any scalar field f(u) (u∈Σ),˜f(u,θ) :=f(u)+θµf(1)
µ(u)+1
2θµθνf(2)
µνis called
as the super field, where f(1)andf(2)represent a one-form field and a two-form field, respectively.
By using the super fields, the gauge fixed action Sgfcan be rewritten as
Sgf=/integraldisplay
Σ/integraldisplay
d2θ/bracketleftbigg
ηiDXi−λid∗HAi+1
2αij(X)ηiηj/bracketrightbigg
, (169)
whereD:=θµ∂
∂uµ. This is the final result in this section. Hereafter, we write S0
gf:=/integraltext
Σ/bracketleftbig
ηiDXi−λid∗HAi/bracketrightbig
and
S1
gf:=/integraltext
Σαijηiηj/2.
D. Equivalence between deformation quantization and topol ogical string theory
Wereturntothediscussionaboutthedeformationquantization. H ereweseethatthe equivalenceofthedeformation
quantization and the topological string theory, and introduce the perturbation theory of the topological string theory,
which is equal to Kontsevich’s deformation quantization [17].
1. Path integral as L∞map
Here we summarize correspondence between Path integral with L∞map.
First we note that the map: α:=α(x)µνηµην/2/ma√sto→Sα:=S1
gf=/integraltext
Σα(x)µνηµην/2 is isomorphic, because
{Sα1,Sα2}BV=S{α1,α2}BV.
SD operator satisfies the conditions of codifferential operator QinL∞algebra, where the vector space and the
degree of the space correspond to C(Ψ) and the ghost number, respectively.
The path integral/integraltext
ei
/planckover2pi1S0
gfgives the deformation quantization F0+F1. The master equation
Qei
/planckover2pi1Sα= 0 (170)
withQ=△SDis corresponding to the L∞map’s condition QF= 0
Forαr:=αi1,i2,···,imηi1ηi2···ηim/m! with a positive integer m,Fn:V⊗n
1⊗→V2is given by
Fn(α1,...,α n)(f1⊗···⊗fm)(x) :=/integraldisplay
ei
/planckover2pi1S0i
/planckover2pi1Sα1···i
/planckover2pi1SαnO(f1,...,f m), (171)
whereSαris the expansion of Sα, and is defined as
Sαr:=/parenleftbigg/integraldisplay
Σ1
m!αi1···im(X)ηi1···ηim/parenrightbigg/vextendsingle/vextendsingle/vextendsingle/vextendsingle
Φ∗=∂ϕ, (172)
andOis chosen to satisfy
Q◦F=F◦Q, (173)
where F:=F0+F1+F2+···. We putOas follow:
O(f1,...,f m) =/integraldisplay
Bm[X(t1,θ1))···fm(X(tm,θm))](m−2)δx(X(∞)), (174)
≡/integraldisplay
1=t1>t2>···>tm=0f1(ψ(1))m−1/productdisplay
k=2∂ik/bracketleftbig
f(ψ(tk))A+ik(tk)/bracketrightbig
fm(ψ(0))δx(ψ(∞)) (175)26
where the subscript ( m−2) denotes that ( m−2) forms are picked up from the products of super fields, and Bm
represents the surface of the disk Σ, i.e., tis the parameter specifying the position on the boundary ∂Σ (1 =t1>
t2>···>tm−1>tm= 0).
To be exact, the action and fields include gauge fixing terms, ghost fi elds and anti-field. Finally, the deformation
quantization is given as follow:
(f ⋆g)(x) =/integraldisplay
DΦf(ψ(1))g(ψ(0))δ(xi−ψi(∞))ei
/planckover2pi1Sgf. (176)
2. Perturbation theory
Nowweseethattheperturbationtheoryofthetopologicalstring theory. First, wewritetheactionas Sgf=S0
gf+S1
gf.
The first term is defined as
S0
gf=/integraldisplay
Σ/bracketleftbig
Ai∧(dξi+∗Hdλi)+cid∗Hdγi/bracketrightbig
, (177)
whereξi≡ψi−xi, and we have expanded ψiaroundxi. The path integral of an observable quantity /an}bracketle{tO/an}bracketri}htis given by
/integraldisplay
ei
/planckover2pi1SgfO=∞/summationdisplay
n=0in
/planckover2pi1nn!/integraldisplay
ei
/planckover2pi1S0
gf(S1
gf)nO, (178)
where/integraltext
:=/integraltext
DξDADcDγDλ. This expansion corresponds to the summation of all diagrams by th e contractions of
all pairs in terms of fields and ghost fields. From equation (177), pro pagators are inverses of
d⊕∗Hd, d∗Hd. (179)
Here we assume that the disk is the upper complex plane: Σ = {z|z=u+iv, u,v∈R, v≥0}withi2=−1, and
the boundary is ∂Σ ={z|z=u, u∈R}. (Rrepresents the real number space, and zdenotes a complex number.)
The Hodge operator ∗His defined by
/braceleftbigg
∗Hdu=dv
∗Hdv=−du/ma√sto−→/braceleftbigg
∗Hdz=−idz
∗Hdz=idz, (180)
wherezrepresents the complex conjugate of z. Moreover,
dz=du∂
∂u+dv∂
∂v=dz∂
∂z+dz∂
∂z, (181)
δz(w) :=δ(w−z)duw∧dvw,/integraldisplay
δz(w) = 1, (182)
wherew∈Cwith the complex number plane C, andw≡uw+ivw.
Now, we calculate Green functions of d⊕∗Hdandd∗Hd, because the Green functions are inversesof these operators:
DwG(z,w) = i/planckover2pi1δz(w), (183)
whereDw=dw⊕∗Hdwordw∗Hdw. The solution depends on the boundary condition. In the case that zandw
satisfy the Neumann boundary condition, a solution is a function of
φh(z,w) :=1
2ilog(z−w)(z−w)
(z−w)(z−w). (184)
On the other hand, zandwsatisfy the Dirichlet boundary condition, a solution is a function of
ψh(z,w) := log/vextendsingle/vextendsingle/vextendsingle/vextendsinglez−w
z−w/vextendsingle/vextendsingle/vextendsingle/vextendsingle. (185)
TheNeumannboundaryconditionis0 = ∂u1G(z,w)|u2=0,andtheDirichletboundaryconditionis0 = ∂u2G(z,w)|u1=0.27
The propagators are given by
/an}bracketle{tγk(w)cj(z)/an}bracketri}ht=i/planckover2pi1
2πδk
jψh(z,w), (186)
/an}bracketle{tξk(w)Aj(z)/an}bracketri}ht=i/planckover2pi1
2πδk
jdzφh(z,w), (187)
/an}bracketle{t(∗Hdγk)(w)cj(z)/an}bracketri}ht=i/planckover2pi1
2πδk
jδwφh(z,w), (188)
and so on. From these propagators, we can obtain diagram rules co rresponding to the deformation quantization. In
section IVE, we will introduce exact diagram rules.
To obtain the star product, we choice
Ox=f(X(1))g(X(0))δx(ψ(∞)) (189)
E. Diagram rules of deformation quantization
From the perturbation theory of the topological string theory, w e can obtain the following diagram rules of the star
product, which is first given by Kontsevich [17, 30]:
(f ⋆g)(x) =f(x)g(x)+∞/summationdisplay
n=1/parenleftbiggi/planckover2pi1
2/parenrightbiggn/summationdisplay
Γ∈GnwΓBΓ,α(f,g). (190)
where Γ,BΓ,α(f,g) andwΓare defined as follows:
Definition. 1 Gnis a set of the graphs Γwhich have n+ 2vertices and 2nedges. Vertices are labeled by symbols
“1”, “2”,..., “n”, “L”, and “R”. Edges are labeled by symbol (k,v), wherek= 1,2,...,n,v= 1,2,...,n,L,R , and
k/ne}ationslash=v.(k,v)represents the edge which starts at “ k” and ends at “ v”. There are two edges starting from each vertex
withk= 1,2,...,n;LandRare the exception, i.e., they act only as the end points of the edges. Hereafter, VΓand
EΓrepresent the set of the vertices and the edges, respectivel y.
Definition. 2 BΓ,α(f,g)is the operator defined by:
BΓ,α(f,g) :=/summationdisplay
I:EΓ→{i1,i2,···,i2n}
n/productdisplay
k=1
/productdisplay
e∈EΓ,e=(k,∗)∂I(e)
αI((k,v1
k),(k,v2
k))
×
/productdisplay
e∈EΓ,e=(∗,L)∂I(e)
f
×
/productdisplay
e∈EΓ,e=(∗,R)∂I(e)
g
, (191)
where,Iis a map from the list of edges ((k,v1,2
k)),k= 1,2,...,nto integer numbers {i1,i2,···,i2n}. Here1≤in≤d;
drepresents a dimension of the manifold M.BΓ,α(f,g)corresponds to the graph Γin the following way: The vertices
“1”, “2”,..., “n”, correspond to the Poisson structure αij.RandLcorrespond to the functions fandg, respectively.
The edgee= (k,v)represents the differential operator ∂(iorj)acting on the vertex v.
The simplest diagram for n= 1is shown in Fig. 4(a), which corresponds to the Poisson bracket: {f,g}=/summationtext
i1,i2αi1i2(∂xi1f)(∂xi2g). The higher order terms are the generalizations of this Pois son bracket.
Figure4(b)shows a graph Γex.2withn= 2corresponding to the list of edges
((1,L),(1,R),(2,R),(2,3)); (192)
in addition, the operator BΓex.2,αis given by
(f,g)/ma√sto→/summationdisplay
i1,···,i4(∂xi3αi1i2)αi3i4(∂xi1f)(∂xi2∂xi4g). (193)
Definition. 3 We put the coordinates for the vertices in the upper-half com plex planeH+:={z∈C|Im(z)>0}
(Crepresents the complex plain; Im(z)denotes the imaginary part of z). Therefore, RandLare put at 0and1,
respectively. We associate a weight wΓwith each graph Γ∈Gnas
wΓ:=1
n!(2π)2n/integraldisplay
Hnn/logicalanddisplay
k=1/parenleftig
dφh
(k,v1
k)∧dφh
(k,v2
k)/parenrightig
, (194)28
FIG. 4. (a): The graph Γ ex.1∈G1corresponding to Poisson bracket. (b): A graph Γ ∈G2correspond to the list of edges:
((1,L),(1,R),(2,R),(2,1))/mapsto→ {i1,i2,i3,i4}.
whereφis defined by
φh
(k,v):=1
2iLog/parenleftbigg(q−p)(¯q−p)
(q−¯p)(¯q−¯p)/parenrightbigg
. (195)
pandqare the coordinates of the vertexes “ k” and “v”, respectively. ¯prepresents the complex conjugate of p∈C.
Hndenotes the space of configurations of nnumbered pair-wise distinct points on H+:
Hn:={(p1,···,pn)|pk∈H+, pk/ne}ationslash=plfork/ne}ationslash=l}. (196)
Here we assume that H+has the metric:
ds2= (d(Re(p))2+d(Im(p))2)/(Im(p))2, (197)
withp∈H+;φh(p,q)is the angle which is defined by (p,q)and(∞,p), i.e.,φh(p,q) =∠pq∞with the metric (197).
For example, wΓex.1corresponding to Fig. 4(a)is calculated as:
wΓex.1=2
1!(2π)2/integraldisplay
H1d1
2iLog/parenleftbiggp2
p2/parenrightbigg
∧d1
2iLog/parenleftbigg(1−p)2
(1−p)2/parenrightbigg
= 1, (198)
where we have included the factor “ 2” arising from the interchange between two edges in Γ.wΓcorresponding to the
Fig.4(b)is
wΓ(b)=1
2!(2π)4/integraldisplay
H2d1
iLog/parenleftbiggp1
p1/parenrightbigg
∧d1
iLog/parenleftbigg1−p1
1−p1/parenrightbigg
∧d1
iLog/parenleftbiggp2
p2/parenrightbigg
∧d1
2iLog/parenleftbigg(p1−p2)(p1−p2)
(p1−p2)(p1−p2)/parenrightbigg
=1
2!(2π)4/integraldisplay
H2d1
iLog/parenleftbiggp1
p1/parenrightbigg
∧d1
iLog/parenleftbigg1−p1
1−p1/parenrightbigg
∧d(2arg(p2))∧d|p2|∂
∂|p2|1
2iLog/parenleftbigg(p1−p2)(p1−p2)
(p1−p2)(p1−p2)/parenrightbigg
=1
2!(2π)4/integraldisplay
H2d1
iLog/parenleftbiggp1
p1/parenrightbigg
∧d1
iLog/parenleftbigg1−p1
1−p1/parenrightbigg
∧d1
iLog/parenleftbiggp2
p2/parenrightbigg
∧d1
iLog/parenleftbigg1−p2
1−p2/parenrightbigg
=w2
1
2!
=1
2, (199)
wherep1andp2are the coordinates of vertexes “ 1” and “2”, respectively. Here, we have used the following facts:
/integraldisplay∞
0d|p2|∂|p2|Log/parenleftbigg(p1−p2)(p1−p2)
(p1−p2)(p1−p2)/parenrightbigg
= lim
Λ→∞Log/parenleftbigg(p1−Λeiarg(p2))(p1−Λeiarg(p2))
(p1−Λe−iarg(p2))(p1−Λe−iarg(p2))/parenrightbigg
= lim
Λ→∞Log/parenleftbigg(1−Λeiarg(p2))(1−Λeiarg(p2))
(1−Λe−iarg(p2))(1−Λe−iarg(p2))/parenrightbigg
,
/integraldisplay
|p1|>ΛdLog/parenleftbiggp1
p1/parenrightbigg
∧dLog/parenleftbigg1−p1
1−p1/parenrightbigg
Λ→∞−→/integraldisplay
|p1|>ΛdLog/parenleftbiggp1
p1/parenrightbigg
∧dLog/parenleftbiggp1
p1/parenrightbigg
= 0. (200)
Generally speaking, the integrals are entangled for n≥3graphs, and the weight of these are not so easy to evaluate
as Eq.(199).
Note that the above diagram rules also define the twisted element as the following relation: ( f⋆g)≡µ◦F(f⊗g).29
FIG. 5. A four vertexes graph, where the white circle and the w hite square represent αAandαF, respectively; the dotted
arrow, waved arrow, and real arrow represent ∂p,∂s, and∂X, respectively.
F. Gauge invariant star product
From Eq. (77), the Poisson structure corresponding to our mode l is
αij=
0ηµν0
−ηµν−qˆFµν−qǫabcsaAb
µ
0qǫabcsaAb
µǫabcsc
, (201)
where the symbols iandjrepresent indexes of the phase space ( TX,ωp,s). We separate the Poisson structure as
follows:
α:=
0ηµν0
−ηµν0 0
0 0 0
+
0 0 0
0 0−qǫabcAb
µsa
0qǫabcsaAb
µǫabcsa
+
0 0 0
0qˆFµν0
0 0 0
≡α0+αA+αF. (202)
Here, forf=f0+faσa,∂saf:=fa(a=x,y,z), wheref0,x,y,zare functions Xandp. Becauseα0is constant and
αAandαFare functions of Xµands, and any function fis written as f=f0+/summationtext
a=x,y,zfasa(f0,aonly depends on
Xandp), then we obtain additional diagram rules:
A1. Two edges starting from αFconnect with both vertices “ L” and “R”.
A2. At least one edge from vertices α0orαFconnect with vertices “ L” or “R”.
A3. A number of the edges entering αAis one or zero.
We also separate the graph Γ into Γ α0, ΓαAand Γ αF. Here, we define the numbers of vertices α0,αA, andαF
asnα0,nαA, andnαF, respectively. Γ αFis the graph consisted by vertices corresponding to αF, and “L” and “R”,
and edges starting from these vertices. We consider Γ αFas a cluster, and define Γ αAas the graph consisted by the
vertices corresponding to αA, which acts on the cluster corresponding to Γ αF. Γα0is the rest of the graph Γ without
ΓαAand Γ αF. Here, we label vertexes Γ αF, ΓαAand Γ α0by “k= 1−nαF”, “k= (nαF+ 1)−(nαF+nαA)” and
“k= (nαF+nαA+1)−(nαF+nαA+nα0)”, respectively. The edge starting from “ k” and ending to “ v1,2
k” represents
(k,v1,2
k).
Next, we calculate weight wnαFand the operator BΓαF,αFcorresponding to Γ αF, and later those for Γ αAor0.
Separation of graph Γ
We now sketch the proof of wΓBΓ,α=wnα0BΓα0,α0·wnαABΓαA,αA·wnαFBΓαF,αF, wherewnαa=wnαa
1
nαa!fora=
0,A,F, andw1is given by Eq. (198).
From the additional rule A1, each operator corresponding to vert exesαFand edges ( αF,LorR) acts onfandg
independently. Thus wαF∼wnαF
1. Secondly we consider the graph which consists of four vertexes c orresponding to
αA,αF, and “L”and“R” as shownin Fig. 5. We alsoassume that one edge ofthe vertex corr espondingto αAconnects
with a vertex corresponding to αF. In this case, from additional diagram rule A3, another edge of the vertex has to
connect with “ L” or “R”. Since we can exchange the role “ R” and “L” by the variable transformation p/ma√sto→1−p,
(p∈H+), we assume that one edge of the vertex corresponding to αAconnect with “ L”. The weight wΓin this case is30
FIG. 6. This figure shows the calculation method of the graph ( a), where the dotted arrow and real arrow represent the
derivative with respect to pandX, respectively, and the white circle and the white triangle r epresent α0andαF, respectively.
We rewrite the graph ( a) as the graph ( c) which is given by the cluster represented by the big circle a nd the operators into it,
where the big circle represents the graph ( b).
given by Eq. (199), i.e., the integrals for the weight is given by replacin g coordinate of the vertex corresponding to αF
with coordinate of “ R” inH+. This result can be expanded to every graph though a graph include s the vertices α0.
For example, we illustrate the calculation of a six vertices graph, whic h only includes α0andαF, in Fig. 6. At first
we make the cluster having only vertices αF,fandg(fig. 6(b)), which is corresponding to the following operator:
wnαF
nαF!αi1i2
Fαi3i4
F(∂pi1∂pi3f)(∂pi2∂pi4g). The edges from the vertices act on the cluster independently (fi g. 6(c)); we
obtain the following operator:wnα0
nα0!wnαF
nαF!αj1j2
0αj3j4
0(∂Xj1αi1i2
F)(∂Xj3αi3i4
F)(∂Xj2∂pi1∂pi3f)(∂Xj4∂pi2∂pi4g).
The position of each vertex corresponding to αAandαFcan be move independently in integrals, and the entangled
integral does not appear. Therefore the weight wnαAof a graph Γ αA∼wnαA
1only depends on the number of vertexes
correspondingto αAandαF, andwΓ=wnαA·wnαFholds generally. From additional rule A2, we can similarly discuss
aboutagraphΓ α0, andobtain wnα0∼wnα0
1. Finally,wecancountthecombinationof nα0,nαAandnαF, anditisgiven
by(nα0+nαA+nαF)!
(nαA+nαF)!nα0!·(nαA+nαF)!
nαA!nαF!. Therefore we obtain the Eq: wΓBΓ,α=wnα0BΓα0,α0·wnαABΓαA,αA·wnαFBΓαF,αF.
The summation of each graph is easy, and we can derive the star pro duct:f ⋆g=µ◦FA(f⊗g), where twisted
elementFAis written as follow:
FA= exp/braceleftbiggi/planckover2pi1
2/parenleftbig
∂Xµ⊗∂pµ−∂pµ⊗∂Xµ/parenrightbig/bracerightbigg
◦exp/braceleftbiggi/planckover2pi1
2εijksk/parenleftbig
Ai
µ∂pµ⊗∂sj+∂si⊗∂sj−∂sj⊗Ai
µ⊗∂pµ/parenrightbig/bracerightbigg
◦exp/braceleftbiggi/planckover2pi1
2/parenleftbig
Fa
µνsa+Fµν/parenrightbig
∂pµ⊗∂pν/bracerightbigg
. (203)
Because the action IAincluding a global U(1) ×SU(2) gauge field, the action IAis written as IA=FA◦F−1
0I0,
thus, the mapF0/mapsto→A:I0/ma√sto→IAis given byF0/mapsto→A=FA◦F−1
0, i.e.,
F(0/mapsto→A)= exp/braceleftbiggi/planckover2pi1
2/parenleftbig
∂Xµ⊗∂pµ−∂pµ⊗∂Xµ/parenrightbig/bracerightbigg
◦exp/braceleftbiggi/planckover2pi1
2εijksk/parenleftbig
Ai
µ∂pµ⊗∂sj+∂si⊗∂sj−∂sj⊗Ai
µ⊗∂pµ/parenrightbig/bracerightbigg
◦exp/braceleftbiggi/planckover2pi1
2/parenleftbig
Fa
µνsa+Fµν/parenrightbig
∂pµ⊗∂pν/bracerightbigg
◦exp/braceleftbigg
−i/planckover2pi1
2εijksk∂si⊗∂sj/bracerightbigg
◦exp/braceleftbigg
−i/planckover2pi1
2/parenleftbig
∂Xµ⊗∂pµ−∂Xµ⊗∂pµ/parenrightbig/bracerightbigg
. (204)
The inverse map is given by the replacement of i by −i in the map (204).31
V. TWISTED SPIN
In this section, we will derive the twisted spin density, which corresp onds to the spin density in commutative
spacetime without the background SU(2) gauge field.
First, we will derive a general form of the twisted spin current in sec tion VA, which is written by using the twisted
variation operator. This operator is constituted of the coproduc t and twisted element; the coproduct reflects the
action rule of the global SU(2) gauge symmetry generator, and th e twisted element represents gauge structure of the
background gauge fields.
In section VB, at first, we will calculate the twisted spin density of th e so-called Rashba-Dresselhaus model in
the Wigner representation using the general form of the twisted s pin density, and next, we will find the twisted spin
operator in real spacetime using correspondence between opera tors in commutative spacetime and noncommutative
phasespace.
A. Derivation of a twisted spin in Wigner space
The Lagrangian density in the Wigner space is given by
L(X,p) =/parenleftbigg
p0−p2
2m/parenrightbigg
⋆/parenleftbig
ψψ†/parenrightbig
, (205)
wheremis the electric mass, f ⋆g:=µ◦FA(f⊗g) for any functions fandg.
The variation corresponding to the infinitesimal global SU(2) gauge transformation is defined as
δsaψ= iϑsaψ, (206)
δsaψ†=−iψ†sa, (207)
δsaxµ= 0, (208)
whereθrepresents an infinitesimal parameter. Therefore, the variation of the Lagrangian density L(X,p) :=ˆL⋆ψψ†
is given by
δsa(L(X,p)) :=ˆL⋆iθsa⋆ψψ†−ˆL⋆ψψ†⋆iθsa(209)
Here we introduce the Grassmann numbers λ1,2,3(λ3:=λ1λ2), the product µ, and the coproduct △η, where the
coproduct satisfies
△η(f) :=f⊗η+η⊗f (210)
for any functions fand operator η. The equation (209) can be rewritten as
δsa(L) =/integraldisplay
dλ3µ◦FA/parenleftig
ˆL⊗/parenleftbig
iλ1θsa⋆λ2ψψ†+λ2ψψ†⋆iλ1θsa/parenrightbig/parenrightig
=/integraldisplay
dλ3µ◦FA/parenleftig
ˆL⊗µ◦FA△iλ1θsa/parenrightig
=/integraldisplay
dλ3µ◦FA(id⊗µ)◦(id⊗FA)◦(id⊗△iλ1θsa)◦(ˆL⊗λ2ψψ†)
= i/integraldisplay
dλ3µ◦(id⊗µ)◦(id⊗△)FA◦(id⊗FA)
◦(id◦θ1/2⊗θ1/2)◦(id⊗△λ1sa)◦(ˆL⊗λ2ψψ†), (211)
where we used/integraltextdλiλj=δij(i,j= 1,2,3) with the Kronecker delta δijHere we introduce the following symbols:
ˆµ:=µ◦(id⊗µ), (212)
ˆFA:= (id⊗△)FA◦(id⊗FA), (213)
ˆθ:=θ1/2⊗θ1/2, (214)
ˆ△λ1sa:= (id⊗△λ1sa), (215)32
where the coproduct in the differential operator space is defined a s
△(dn) :=/summationdisplay
i+j=n
i≥0, j≥0di⊗dj. (216)
Here vectors{d0,d1,...}corresponding to following operators: d0:= id anddn:= (1/n!)(∂n/∂pn), ord0:= id and
dn:= (1/n!)(∂n/∂xn) (n= 1,2,...), anddlare bases of a vector space B(k) :=/circleplustext∞
l=0kdl(krepresents a scalar). The
coproduct△in the vector space B(k) satisfies the coassociation law: ( △⊗id)⊗△= (id⊗△)◦△because
(△⊗id)◦△(dn) =/summationdisplay
i+j=n
i≥0, j≥0△(di)⊗dj
=/summationdisplay
i+j=n
i≥0, j≥0
/summationdisplay
k+l=i
k≥0, l≥0dk⊗dl⊗dj
=/summationdisplay
k+l+j=n
k≥0, l≥0, j≥0dk⊗dl⊗dj (217)
and
(id⊗△)◦△(dn) =/summationdisplay
i+j=n
i≥0, j≥0di⊗△(dj)
=/summationdisplay
i+k+l=n
i≥0, k≥0, l≥0di⊗dk⊗dl. (218)
It represents the Leibniz rule with respect to the differential oper ator∂µ. For instance, (id ⊗△)(∂µ⊗∂ν) :=
∂µ⊗△(∂ν) =∂µ⊗∂ν⊗id+∂µ⊗id⊗∂ν; it corresponds to the following calculation:
∂µf·∂ν(g·h) =∂µf·∂νg·h+∂µf·g·∂νh. (219)
The variation (211) is rewritten as
δsaL(X,p) = i/integraldisplay
dλ3ˆµ◦ˆFA◦ˆθ◦ˆ△λ1sa◦(ˆL⊗λ2ψψ†). (220)
If we replace ˆ△t
λ1sa:=ˆF−1
Aˆ△λ1saF0withˆ△λ1sain equation (220), the integrals in terms of xandpof the
right-hand side of Eq. (220) become zero because ˆFA◦ˆ△t
λ1sadoes not include the SU(2) field, which breaks
the global SU(2) gauge symmetry, in the case that the parameter θis constant. Therefore, for the action S:=/integraltext
dDimXdDimpL(X,p)/(2π/planckover2pi1)Dim,
δt
saS:= tr/integraldisplay/integraldisplay/integraldisplay
dλ3dDimXdDimp
(2π/planckover2pi1)Dimˆµ◦ˆFA◦ˆθ◦ˆ△t
λ1sa(ˆL⊗λ2ψψ†) (221)
is the infinitesimal SU(2) gauge transformation with background SU (2) gauge fields.
Becauseδt
saS= 0, we can write
δt
saS=/integraldisplay
dDimXθ/parenleftbig
∂µjt
µ/parenrightbig
. (222)
In the case that the infinitesimal parameter depends on the space time coordinate, this equation can be written as
δt
saS=/integraldisplay
dDimXθ(X)/parenleftbig
∂µjt
µ/parenrightbig
=−/integraldisplay
dDimX/parenleftbigg∂θ(X′)
∂X′µ/parenrightbigg
jt
µ(X′). (223)33
Therefore, we obtain the twisted Noether current
jt
µ=−δt
saS
δ(∂µθ(X)). (224)
In particular, the twisted spin
St
a=/integraldisplay
dXjt
0
=/integraldisplay
dXδt
saS
δ(∂Tθ)(225)
is conserved quantity. Here, we assumed that the SU(2) gauge is s tatic one. However, we do not use this condition in
the derivation of the twisted Noether charge and current density . Then, we can derive the virtual twisted spin with a
time-dependent SU(2) gauge: ˜St
a. In this case, we only use the time-dependent SU(2) gauge field str engthFa
µν, which
has non-zero space-time components Fa
0i(i= 1,2,...,Dim−1).
Here, we discuss the adiabaticity of the twisted spin. In the case th at SU(2) gauge fields have time dependence, the
twisted spin is not conserved. Now, we assume that Aa
µ=λ(t)Ca
µ(a=x,y,z) with constant fields Ca
µ= (0,Ca);λ(t)
is an adequate slowly function dependent on time. Because ˜St
aincludesF−1
µν∼1
(1+(˙λ)2C·C)/parenleftbigg
0 ˙λC
−˙λCλ−2[Ci,Cj]−1/parenrightbigg
with˙λ≡dλ/dt, the difference between ˜St
aandSt
acomes from only that between inverse of field strength: ∆ F−1∼
1
1+(˙λ)2λ−2[Ci,Cj]−1−λ−2[Ci,Cj]−1∼(˙λ/λ)2[Ci,Cj]−1. Therefore we obtain
dSt
a
dt=O(˙λ2). (226)
This means that St
ais the adiabatic invariance. Namely, for the infinitely slow change in λ(t) during the time period
T(→∞),St
aremains constant while ∆ λ=λ(T)−λ(0) is finite. This fact is essential for the spin-orbit echo proposed
in [11].
B. Rashba-Dresselhaus model
Here we apply the formalism developed so far to an explicit model, i.e., th e so-called Rashba-Dresselhaus model
given by
H=ˆp2
2m+α(ˆpxˆσy−ˆpyˆσx)+β(ˆpxˆσx−ˆpyˆσy)+V(ˆx) (227)
with a potential V(ˆx), whereαandβare the Rashba and Dresselhaus parameters, respectively. Comp leting square
in terms of ˆ p, we obtain Ax
x=−2mβ/(/planckover2pi1q),Ay
x=−2mα/(/planckover2pi1q),Ax
y= 2mα/(/planckover2pi1q),Ay
y= 2mβ/(/planckover2pi1q),A0=m(α2+β2)/e,
andAz
x,y=Ax,y
z=Az
z=Ax,y,z
0=Ax,y,z= 0, where q=|e|/(mc2).
To calculate the twisted symmetry generator ˆ△t(λ2sa), we first consider the ˆF0. (id⊗△)F0is given by
(id⊗△)F0= exp/braceleftbiggi/planckover2pi1
2/parenleftig
∂Xµ⊗1⊗∂pµ+∂Xµ⊗∂pµ⊗1−∂pµ⊗1⊗∂Xµ−∂pµ⊗∂Xµ⊗1/parenrightig
+i
2εijk/parenleftig
sk∂si⊗1⊗∂sj+sk∂si⊗∂sj⊗1/parenrightig/bracerightbigg
, (228)34
andˆF0is given by
ˆF0= exp/braceleftbiggi/planckover2pi1
2/parenleftig
∂Xµ⊗1⊗∂pµ+∂Xµ⊗∂pµ⊗1−∂pµ⊗1⊗∂Xµ−∂pµ⊗∂Xµ⊗1/parenrightig
+i
2εijk/parenleftig
sk∂si⊗1⊗∂sj+sk∂si⊗∂sj⊗1/parenrightig/bracerightbigg
◦exp/braceleftbiggi/planckover2pi1
2/parenleftig
1⊗∂Xµ⊗∂pµ−1⊗∂pµ⊗∂Xµ/parenrightig
+i
2εijk/parenleftig
1⊗sk∂si⊗∂sj/parenrightig/bracerightbigg
= exp/braceleftbiggi/planckover2pi1
2/parenleftig
1⊗∂Xµ⊗∂pµ+∂Xµ⊗1⊗∂pµ+∂Xµ⊗∂pµ⊗1
−1⊗∂pµ⊗∂Xµ−∂pµ⊗1⊗∂Xµ−∂pµ⊗∂Xµ⊗1/parenrightig
+i
2εijk/parenleftig
1⊗sk∂si⊗∂sj+sk∂si⊗1⊗∂sj+sk∂si⊗∂sj⊗1/parenrightig/bracerightbigg
≡G0. (229)
Similarly, ˆFAis given by
ˆFA= exp/braceleftbiggi/planckover2pi1
2/parenleftig
1⊗∂Xµ⊗∂pµ+∂Xµ⊗1⊗∂pµ+∂Xµ⊗∂pµ⊗1
−1⊗∂pµ⊗∂Xµ−∂pµ⊗1⊗∂Xµ−∂pµ⊗∂Xµ⊗1/parenrightig/bracerightig
◦exp/braceleftbiggi
2εijk/parenleftig
1⊗sk∂si⊗∂sj+sk∂si⊗1⊗∂sj+sk∂si⊗∂sj⊗1
+1⊗skAi
µ∂pµ⊗∂sj+skAi
µ∂pµ⊗1⊗∂sj+skAi
µ∂pµ⊗∂sj⊗1
−1⊗∂sj⊗skAi
µ∂pµ−∂sj⊗1⊗skAi
µ∂pµ−∂sj⊗skAi
µ∂pµ⊗1/parenrightig/bracerightig
◦exp/braceleftbiggi/planckover2pi1
2/parenleftig
1⊗ˆFµν∂pµ⊗∂pν+ˆFµν∂pµ⊗1⊗∂pν+ˆFµν∂pµ⊗∂pν⊗1/parenrightig/bracerightbigg
≡GA
Xp◦GA
sp◦GA
pp. (230)
We note that the operators G0andGA
Xp,sp,pphave each inverse operator, which are denoted by G0andGA
Xp,sp,pp,
respectively. Here, the overline −represents the complex conjugate.
Because the twisted variation is ˆ µ◦ˆFA◦ˆθ◦ˆF−1
A◦ˆ△λ1sa◦F0(L⊗λ2ψψ†), the infinitesimal parameter ˆθbecomes
an operator ˆFA◦ˆθ◦ˆF−1
A. It is calculated by using the operator formula
eBCe−B=∞/summationdisplay
n=01
n![B,[B,···[B,/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright
nC]···]] (231)
for any operators BandC. In the calculation, one will use the following formula in midstream:
/summationdisplay
l=0Dl
(l+1)!=/integraldisplay1
0dλeλD, (232)
/summationdisplay
l=0Dl
(l+2)!=/integraldisplay1
0dλ/integraldisplayλ
0dλ′eλ′D, (233)
and so on, for any operator D.
From these results of the calculations, we obtain the twisted spin as follow
st
a=µ◦FA◦(sa⊗id)◦˜Υ◦(id⊗G)◦˜Υ†, (234)35
where˜Υ is defined as
˜Υ :=1
2/bracketleftbigg
ei
2(α+β)σ−/parenleftbig
∂px+∂py/parenrightbig
⊗ei
2(α−β)σ+/parenleftbig
∂px−∂py/parenrightbig/bracketrightbigg
◦/bracketleftigg
e−i
2/parenleftbig
∂X⊗∂Y+∂Y⊗∂X/parenrightig
◦/parenleftbig
1
2m(α2−β2)⊗σz/parenrightbig
+e−i
2/parenleftbig
∂X⊗∂Y+∂Y⊗∂X/parenrightig
◦/parenleftbig
σz⊗1
2m(α2−β2)/parenrightbig/bracketrightigg
+ei
2(α+β)σ−/parenleftbig
∂px+∂py/parenrightbig
⊗sin2/parenleftigg
i/planckover2pi1/parenleftbig
∂X−∂Y/parenrightbig
8√
2m(α+β)/parenrightigg
,(235)
whereσ±:=σx±σy.
Finally, we will rewrite the twisted spin as an operator form in commuta tive spacetime. Roughly speaking, the
operator in the commutative spacetime. and the one in the noncomm utative Wigner space have the following rela-
tions (the left-hand side represents operators on the Wigner spa ce; the right-hand side represents operators on the
commutative spacetime):
Xµ⋆⇔ˆxµ, (236)
pµ⋆⇔ˆpµ, (237)
sa⋆⇔ˆsa, (238)
i/planckover2pi1∂pµ⋆⇔ˆxµ, (239)
−i/planckover2pi1∂Xµ⋆⇔ˆpµ, (240)
because [Xµ,pν]⋆:=Xµ⋆pν−pν⋆Xµ= i/planckover2pi1δµ
νis equal to the commutation of the operator form: [ˆ xµ,ˆpν] = i/planckover2pi1δµ
ν. The
equivalence of st
aand the twisted spin in the operator form on the commutative space time ˆst
acan be confirmed using
the Wigner transformation in terms of ˆ st
a.
The operator form of the twisted spin is given by
st
a=/planckover2pi1
2ψ†Υ†σaΥψ, (241)
where
Υ = lim
x′→x1
2ei
2m(α+β)σ−x+/bracketleftbigg
e−i
2m(α−β)σ+x′
−e−i
2σz
2m2(α2−β2)/parenleftbig← −∂x′∂y+← −∂y′∂x/parenrightbig
+e−i
2σz
2m2(α2−β2)/parenleftbig
∂x′∂y+∂y′∂x/parenrightbig
e−i
2m(α−β)σ+x′
−+2sin2/parenleftbiggi/planckover2pi1(∂x−∂y)
8√
2m(α+β)/parenrightbigg/bracketrightbigg
(242)
withx±:=x±y.
This operator in Eq. (241), when integrated over X, is the conserved quantity for any potential configuration V(ˆx)
as long asα,βare static and the electron-electron interaction is neglected.
VI. CONCLUSIONS
In this paper, we have derived the conservation of the twisted spin and spin current densities. Also the adiabatic
invariant nature of the total twisted spin integrated over the spa ce is shown. Here we remark about the limit of
validity of this conservation law. First, we neglected the dynamics of the electromagnetic field Aµwhich leads to
the electron-electron interaction. This leads to the inelastic electr on scattering, which is not included in the present
analysis, and most likely gives rise to the spin relaxation. This inelastic s cattering causes the energy relaxation and
hence the memory of the spin will be totally lost after the inelastic lifet ime. This situation is analogous to the two
relaxation times T1andT2in spin echo in NMR and ESR. Namely, the phase relaxation time T2is usually much
shorter than the energy relaxationtime T1, and the spin echo is possible for T <T 1. Similar story applies to spin-orbit
echo [11] where the recoveryof the spin moment is possible only within the inelastic lifetime of the spins. However, the
generalization of the present study to the dynamical Aµis a difficult but important issue left for future investigations.
Also the effect of the higher order terms in 1 /(mc2) in the derivation of the effective Lagrangian from Dirac theory
requires to be scrutinized.
Another direction is to explore the twisted conserved quantities in t he non-equilibrium states. Under the static
electric field, the system is usually in the current flowing steady stat e. Usually this situation is described by the
linear response theory, but the far from equilibrium states can in pr inciple be described by the non-commutative
geometry [16, 30]. The nonperturbative effects in this non-equilibriu m states are the challenge for theories, and
deserve the further investigations.36
The author thanks Y.S. Wu, F.C. Zhang, K. Richter, V. Krueckl, J. N itta, and S. Onoda for useful discussions. This
workwassupportedbyPriorityAreaGrants, Grant-in-Aidsunder the Grantnumber21244054,StrategicInternational
Cooperative Program (Joint Research Type) from Japan Science a nd Technology Agency, and by Funding Program
for World-Leading Innovative R and D on Science and Technology (FI RST Program).
[1] M. E. Peshkin and D. V. Schroeder. Introduction to Quantum Field Theory . Addison-Wesley, New York (1995).
[2] J. Fr¨ ohlich and U. M. Studer. Gauge invariance and curre nt algebra in nonrelativistic many-body theory. Rev. Mod. P hys.,
65, 733–802 (1993).
[3] G. Volovik. The Universe in a Helium Droplet . Oxford University Press, Oxford, U.K. (2003).
[4] B. Leurs, Z. Nazario, D. Santiago and J. Zaanen. Non-Abel ian hydrodynamics and the flow of spin in spinorbit coupled
substances. Annals of Physics, 323(4), 907 – 945 (2008).
[5] N. Seiberg and E. Witten. String theory and noncommutati ve geometry. Journal of High Energy Physics, 09, 032 (1999).
[6] M. Chaichian, P. Kulish, K. Nishijima and A. Tureanu. On a Lorentz-invariant interpretation of noncommutative spac e-
time and its implications on noncommutative QFT. Physics Le tters B, 604(1-2), 98–102 (2004).
[7] M. Chaichian, P. Preˇ snajder and A. Tureanu. New Concept of Relativistic Invariance in Noncommutative Space-Time:
Twisted Poincar´ e Symmetry and Its Implications. Phys. Rev . Lett.,94, 151602 (2005).
[8] G. Amelino-Camelia, G. Gubitosi, A. Marcian` o, P. Marti netti, F. Mercati, D. Pranzetti and R. A. Tacchi. First Resul ts
of the Noether Theorem
for Hopf-Algebra Spacetime Symmetries. Progress of Theore tical Physics Supplement, 171, 65–78 (2007).
[9] G. Amelino-Camelia, G. Gubitosi, A. Marcian` o, P. Marti netti and F. Mercati. A no-pure-boost uncertainty principl e from
spacetime noncommutativity. Physics Letters B, 671(2), 298–302 (2009).
[10] S. Maekawa. Concepts in Spin Electronics . OXFORD SCIENCE PUBLICATIONS (2006).
[11] N. Sugimoto and N. Nagaosa. Spin-orbit echo. To apper in Science.
[12] E. L. Hill. Hamilton’s Principle and the Conservation T heorems of Mathematical Physics. Rev. Mod. Phys., 23, 253–260
(1951).
[13] C. Gonera, P. Kosi´ nski, P. Ma´ slanka and S. Giller. Spa ce-time symmetry of noncommutative field theory. Physics Le tters
B,622(1-2), 192–197 (2005).
[14] E. Wigner. On the Quantum Correction For Thermodynamic Equilibrium. Phys. Rev., 40, 749–759 (1932).
[15] J. E. Moyal. Mathematical Proceedings of the Cambridge Philosophical Society. Proc. Cambridge Phil. Soc., 45, 99–124
(1949).
[16] S. Onoda, N. Sugimoto and N. Nagaosa. Theory of Non-Equi libirum States
Driven by Constant Electromagnetic Fields. Progress of The oretical Physics, 116(1), 61–86 (2006).
[17] M. Kontsevich. Deformation Quantization of Poisson Ma nifolds. Letters in Mathematical Physics, 66, 157–216 (2003).
[18] F. Bayen, M. Flato, C. Fronsdal, A. Lichnerowicz and D. S ternheimer. Deformation theory and quantization. I. Defor ma-
tions of symplectic structures. Annals of Physics, 111(1), 61 – 110 (1978).
[19] F. Bayen, M. Flato, C. Fronsdal, A. Lichnerowicz and D. S ternheimer. Deformation theory and quantization. II. Phys ical
applications. Annals of Physics, 111(1), 111 – 151 (1978).
[20] G. Hochschild. On the Cohomology Groups of an Associati ve Algebra. The Annals of Mathematics, 46(1), pp. 58–67
(1945).
[21] D. Happel. Hochschild cohomology of finitedimensional algebras. In M.-P. Malliavin, editor, S´ eminaire d’Alg` ebre Paul
Dubreil et Marie-Paul Malliavin , vol. 1404 of Lecture Notes in Mathematics , pages 108–126. Springer Berlin / Heidelberg.
ISBN 978-3-540-51812-9 (1989).
[22] M. Gerstenhaber. The Cohomology Structure of an Associ ative Ring. The Annals of Mathematics, 78(2), 267–288 (1963).
[23] E. Cartan. Sur la structure des groupes infinis de transf ormation. Annales Scientifiques de l’ ´Ecole Normale Sup´ erieure,
22, 219–308 (0012-9593).
[24] A. S. Cattaneo and G. Felder. A Path Integral Approach to the Kontsevich Quantization Formula. Communications in
Mathematical Physics, 212, 591–611 (2000).
[25] N. Ikeda and I. Ken-iti. Dimensional reduction of nonli near gauge theories. Journal of High Energy Physics, 09, 030
(2004).
[26] I. A. Batalin and G. A. Vilkovisky. Gauge algebra and qua ntization. Physics Letters B, 102(1), 27–31 (1981).
[27] I. A. Batalin and G. A. Vilkovisky. Quantization of gaug e theories with linearly dependent generators. Phys. Rev. D ,28,
2567–2582 (1983).
[28] J. Alfaro and P. Damgaard. Origin of antifields in the Bat alin-Vilkovisky lagrangian formalism. Nuclear Physics B, 404(3),
751–793 (1993).
[29] I. Batalin and G. Vilkovisky. Relativistic S-matrix of dynamical systems with boson and fermion constraints. Phys ics
Letters B, 69(3), 309–312 (1977).
[30] N. Sugimoto, S. Onoda and N. Nagaosa. Gauge Covariant Fo rmulation of the Wigner Representation through Deformatio n
Quantization. Progress of Theoretical Physics, 117(3), 415–429 (2007). |
1304.2766v1.Spin_orbital_liquids_in_non_Kramers_magnet_on_Kagome_lattice.pdf | Spin-orbital liquids in non-Kramers magnet on Kagome lattice
Robert Schaffer1, Subhro Bhattacharjee1;2, and Yong Baek Kim1;3
1Department of Physics and Center for Quantum Materials,
University of Toronto, Toronto, Ontario M5S 1A7, Canada.
2Department of Physics and Astronomy, McMaster University, Hamilton, Ontario L8S 4M1, Canada.
3School of Physics, Korea Institute for Advanced Study, Seoul 130-722, Korea.
(Dated: June 21, 2018)
Localized magnetic moments with crystal-field doublet or pseudo-spin 1/2 may arise in correlated insulators
with even number of electrons and strong spin-orbit coupling. Such a non-Kramers pseudo-spin 1/2 is the
consequence of crystalline symmetries as opposed to the Kramers doublet arising from time-reversal invariance,
and is necessarily a composite of spin and orbital degrees of freedom. We investigate possible spin-orbital
liquids with fermionic spinons for such non-Kramers pseudo-spin 1/2 systems on the Kagome lattice. Using the
projective symmetry group analysis, we find tennew phases that are not allowed in the corresponding Kramers
systems. These new phases are allowed due to unusual action of the time reversal operation on non-Kramers
pseudo-spins. We compute the spin-spin dynamic structure factor that shows characteristic features of these
non-Kramers spin-orbital liquids arising from their unusual coupling to neutrons, which is therefore relevant for
neutron scattering experiments. We also point out possible anomalous broadening of Raman scattering intensity
that may serve as a signature experimental feature for gapless non-Kramers spin-orbital liquids.
PACS numbers:
I. INTRODUCTION
The low energy magnetic degrees of freedom of a Mott
insulator, in the presence of strong spin-orbit coupling, are
described by states with entangled spin and orbital wave
functions.1,2In certain crystalline materials, for ions with
even numbers of electrons, a low energy spin-orbit entan-
gled “pseudo-spin”-1/2 may emerge, which is not protected
by time-reversal symmetry (Kramers degeneracy)12but rather
by the crystal symmetries.3,4Various phases of such non-
Kramers pseudo-spin systems on geometrically frustrated lat-
tices, particularly various quantum paramagnetic phases, are
of much recent theoretical and experimental interest in the
context of a number of rare earth materials including frus-
trated pyrochlores5–9and heavy fermion systems.10,11
In this paper, we explore novel spin-orbital liquids that may
emerge in these systems due to the unusual transformation of
the non-Kramers pseudo-spins under the time reversal trans-
formation. Contrary to Kramers spin-1/2, where the spins
transform as S! Sunder time reversal,12here only one
component of the pseudo-spin operators changes sign under
time reversal:f1;2;3g!f1;2; 3g.3,4This is be-
cause, due to the nature of the wave-function content, the
3component of the pseudo-spin carries a dipolar magnetic
moment while the other two components carry quadrupolar
moments of the underlying electrons. Hence the time rever-
sal operator for the non-Kramers pseudo-spins is given by
T=1K(whereKis the complex conjugation operator),
which allows for new spin-orbital liquid phases. Since the
magnetic degrees of freedom are composed out of wave func-
tions with entangled spin and orbital components, we prefer
to refer the above quantum paramagnetic states as spin-orbital
liquids, rather than spin liquids.
Since the degeneracy of the non-Kramers doublet is pro-
tected by crystal symmetries, the transformation properties of
the pseudo-spin under various lattice symmetries intimatelydepend on the content of the wave-functions that make up
the doublet. To this end, we focus our attention on the ex-
ample of Praseodymium ions (Pr3+) in a localD3denviron-
ment, which is a well known non-Kramers ion that occurs in
a number of materials with interesting properties.5–7Such an
environment typically occurs in Praseodymium pyrochlores
given by the generic formulae Pr 2TM2O7, where TM(= Zr,
Sn, Hf, or Ir) is a transition metal. In these compounds, the
Pr3+ions host a pair of 4 felectrons which form a J= 4
ground state manifold with S= 1andL= 5, as expected due
to Hund’s rules. In terms of this local environment we have
a nine fold degeneracy of the electronic states.4This degener-
acy is broken by the crystalline electric field. The oxygen and
TM ions form a D3dlocal symmetry environment around the
Pr3+ions, splitting the nine fold degeneracy. A standard anal-
ysis of the symmetries of this system (see appendix A) shows
that theJ= 4 manifold splits into three doublets and three
singlets ( j=4= 3Eg+ 2A1g+A2g) out of which one of the
doublets is found to have the lowest energy, usually well sepa-
rated from the other crystal field states.4This doublet (details
in Appendix A), formed out of a linear combination of the
Jz=4withJz=1andJz=2states, is given by
ji=jm=4ijm=1i
jm=2i:(1)
The non-Kramers nature of this doublet is evident from the
nature of the “spin” raising and lowering operators within
the doublet manifold; the projection of the angular momen-
tum raising and lowering operators to the space of doublets
is zero (PJPji= 0 wherePprojects into the doublet
manifold). However, the projection of the Jzoperator to this
manifold is non-zero, and describes the z component of the
pseudo-spin ( 3). In addition, there is a non-trivial projection
of the quadrupole operators fJ;Jzgin this manifold. These
have off-diagonal matrix elements, and are identified with the
pseudo-spin raising and lowering operators ( =1i2).
In a pyrochlore lattice the local D3daxes point to the centre
of the tetrahedra.4On looking at the pyrochlore lattice alongarXiv:1304.2766v1 [cond-mat.str-el] 9 Apr 20132
the [111] direction, it is found to be made out of alternate
layers of Kagome and triangular lattices. For each Kagome
layer (shown in Fig. 1) the local D3daxes make an angle of
cos 1(p
2=3)with the plane of the Kagome layer. We imag-
ine replacing the Pr3+ions from the triangular lattice layer
with non-magnetic ions so as to obtain decoupled Kagome
layers with Pr3+ions on the sites. The resulting structure is
obtained in the same spirit as the now well-known Kagome
compound Herbertsmithite was envisioned. As long as the lo-
cal crystal field has D3dsymmetry, the doublet remains well
defined. A suitable candidate non-magnetic ion may be iso-
valent but non-magnetic La3+. Notice that the most extended
orbitals in both cases are the fifth shell orbitals and the crystal
field at each Pr3+site is mainly determined by the surrounding
oxygens and the transition metal element. Hence, we expect
that the splitting of the non-Kramers doublet due to the above
substitution would be very small and the doublet will remain
well defined. In this work we shall consider such a Kagome
lattice layer and analyze possible Z2spin-orbital liquids, with
gapped or gapless fermionic spinons .
The rest of the paper is organized as follows. In Sec. II, we
begin with a discussion of the symmetries of the non-Kramers
system on a Kagome lattice and write down the most general
pseudo-spin model with pseudo-spin exchange interactions up
to second nearest neighbours. In Sec. III formulate the pro-
jective symmetry group (PSG) analysis for singlet and triplet
decouplings. Using this we demonstrate that the non-Kramers
transformation of our pseudo-spin degrees of freedom under
time reversal leads to a set of ten spin-orbital liquids which
cannot be realized in the Kramers case. In Sec. IV we de-
rive the dynamic spin-spin structure factor for a representa-
tive spin liquid for the case of both Kramers and non-Kramers
doublets, demonstrating that experimentally measurable prop-
erties of these two types of spin-orbital liquids differ qualita-
tively. Finally, in Sec. V, we discuss our results, and propose
an experimental test which can detect a non-Kramers spin-
orbital liquid. The details of various calculations are discussed
in different appendices.
II. SYMMETRIES AND THE PSEUDO-SPIN
HAMILTONIAN
Since the local D3daxes of the three sites in the Kagome
unit cell differ from each other a general pseudo-spin Hamil-
tonian is not symmetric under continuous global pseudo-spin
rotations. However, it is symmetric under various symmetry
transformations of the Kagome lattice as well as time rever-
sal symmetry. Such symmetry transformations play a major
role in the remainder of our analysis. We start by describing
the effect of various lattice symmetry transformations on the
non-Kramers doublet.
We consider the symmetry operations that generate the
space group of the above Kagome lattice. These are (as shown
in Fig. 2(a))
T1,T2: generate the two lattice translations.
=C0
2I: (not to be confused with the pseudo-spin
FIG. 1: A Kagome layer, in the pyrochlore lattice environment. We
consider sites labelled z and z’ replaced by non-magnetic ions, de-
coupling the Kagome layers. The local axis at the u,v and w sites
point towards the center of the tetrahedron on which these lie.
u
vw
FIG. 2: (color online) (a) The symmetries of the Kagome lattice.
Also shown are the labels for the sublattices and the orientation of the
local z-axis. (b) Nearest and next nearest neighbour bonds. Colors
refer to the phases r;r0and0
r;r0, with these being 0 on blue bonds,
1 on green bonds and 2 on red bonds.
operators which come with a superscript) where Iis the
three dimensional inversion operator about a plaquette
center andC0
2refers to a two-fold rotation about a line
joining two opposite sites on the plaquette.
S6=C2
3I: whereC3is the threefold rotation op-
erator about the center of a hexagonal plaquette of the
Kagome lattice.
T=1K: Time reversal.
Here, we consider a three dimensional inversion operator
since the local D3daxes point out of the Kagome plane. The
above symmetries act non-trivially on the pseudo-spin degrees
of freedom, as well as the lattice degrees of freedom. The ac-
tion of the symmetry transformations on the pseudo-spin op-
erators is given by,
S6:f3;+; g!f3;!+;! g;
T:f3;+; g!f 3; ;+g;
C0
2:f3;+; g!f 3; ;+g;
T1:f3;+; g!f3;+; g;
T2:f3;+; g!f3;+; g; (2)
(!= ! 1=ei2
3). Operationally their action on the doublet
(j+i j i )can be written in form of 22matrices. The
translationsT1;T2act trivially on the pseudo-spin degrees of3
freedom, and the remaining operators act as
T=1K; =1; S 6=
!0
0!
; (3)
whereKrefers to complex conjugation. The above expres-
sions can be derived by examining the effect of these operators
on the wave-function describing the doublet (Eq. 1).
We can now write down the most generic pseudo-spin
Hamiltonian allowed by the above lattice symmetries that is
bilinear in pseudo-spin operators. The form of the time-
reversal symmetry restricts our attention to those products
which are formed by a pair of 3operators or those which
mix the pseudo-spin raising and lowering operators. Any
term which mixes 3andchanges sign under the sym-
metry, and can thus be excluded. Under the C3transforma-
tion about a site, the terms C3:3
r3
r0!3
C3(r)3
C3(r0)and
C3:+
r
r0!+
C3(r)
C3(r0). However, the term +
r+
r0(and
its Hermitian conjugate) gain additional phase factors when
transformed; under the C3symmetry transformation, this term
becomesC3:+
r+
r0!!+
C3(r)+
C3(r0). In addition, un-
der thesymmetry, this term transforms as :+
r+
r0!
(r)
(r0). Thus the Hamiltonian with spin-spin exchange
interactions up to next-nearest neighbour is given by
Heff=JnnX
hr;r0i[3
r3
r0+ 2(+
r
r0+h:c:)
+2q(e2ir;r 0
3+
r+
r0+h:c:)]
+JnnnX
hhr;r0ii[3
r3
r0+ 2(0+
r
r0+h:c:)
+2q0(e2i0
r;r 0
3+
r+
r0+h:c:)]; (4)
whereand0take values 0, 1 and 2 depending on the bonds
on which they are defined (Fig. 2(b)).
III. SPINON REPRESENTATION OF THE PSEUDO-SPINS
AND PSG ANALYSIS
Having written down the pseudo-spin Hamiltonian, we now
discuss the possible spin-orbital liquid phases. We do this in
two stages in the following sub-sections.
A. Slave fermion representation and spinon decoupling
In order to understand these phases, we will use the
fermionic slave-particle decomposition of the pseudo-spin op-
erators. At this point, we note that the pseudo-spins satisfy
S= 1=2representations of a “SU(2)” algebra among their
generators (not to be confused with the regular spin rotation
symmetry). We represent the pseudo-spin degrees of freedom
in terms of a fermion bilinear. This is very similar to usual
slave fermion construction for spin liquids13,14. We take
j=1
2fy
j[]fj; (5)where;=";#is defined along the local zaxis andfy(f)
is anS= 1=2fermionic creation (annihilation) operator. Fol-
lowing standard nomenclature, we refer to the f(fy)as the
spinon annihilation (creation) operator, and note that these
satisfy standard fermionic anti-commutation relations. The
above spinon representation, along with the single occupancy
constraint
fy
i"fi"+fy
i#fi#= 1; (6)
form a faithful representation of the pseudo-spin-1/2 Hilbert
space. The above representation of the pseudo-spins, when
used in Eq. 4, leads to a quartic spinon Hamiltonian. Fol-
lowing standard procedure,13,14this is then decomposed us-
ing auxiliary fields into a quadratic spinon Hamiltonian (af-
ter writing down the corresponding Eucledian action). The
mean field description of the phases is then characterized by
the possible saddle point values of the auxiliary fields. There
are eight such auxiliary fields per bond, corresponding to
ij=hfy
ifji;ij=hfi
i2
fji; (7a)
Ea
ij=hfy
i[a]fji;Da
ij=hfi
i2a
fii;
(7b)
wherea(a= 1;2;3) are the Pauli matrices. While Eq. 7a
represents the usual singlet spinon hopping (particle-hole) and
pairing (particle-particle) channels, Eq. 7b represents the cor-
responding triplet decoupling channels. Since the Hamilto-
nian (Eq. 4) does not have pseudo-spin rotation symmetry,
both the singlet and the triplet decouplings are necessary.16,17
From this decoupling, we obtain a mean-field Hamiltonian
which is quadratic in the spinon operators. We write this com-
pactly in the following form17(subject to the constraint Eq.6)
H0=X
ijJij~fy
iUij~fj; (8)
~fy
i=h
fy
i"fi#fy
i# fi"i
; (9)
Uij=
ij ; (10)
=
I; =I
; (11)
whereare the Identity (for = 0) and Pauli matrices
(= 1;2;3) acting on pseudo-spin degrees of freedom, and
represents the same in the gauge space. We immediately
note that
;
= 08;: (12)
The requirement that our H0be Hermitian restricts the coeffi-
cientsijto satisfy
00
ij;ab
ij2=;a0
ij;0b
ij2<: (13)
fora;b2 f 1;2;3g. The relations between ijs and
fij;ij;Eij;Dijgare given in Appendix C.17As a straight
forward extension of the SU(2)gauge theory formulation for4
spin liquids,13,18we find that H0is invariant under the gauge
transformation
~fj!Wj~fj; (14)
Uij!WiUijWy
j; (15)
where theWimatrices are SU(2) matrices of the form Wi=
ei~ ~ ai(~ ( 1; 2; 3)). Noting that the physical pseudo-
spin operators are given by
~ i=1
4~fy
i~~fi; (16)
Eq. 12 shows that the spin operators, as expected, are gauge
invariant. It is useful to define the “ -components” of the Uij
matrices as follows:
Uij=V
ij; (17)
where
V
ij=
ij =J
ij0
0J
ij
; (18)
and
J
ij=0
ij+3
ij1
ij i2
ij
1
ij+i2
ij0
ij 3
ij
: (19)
Under global spin rotations the fermions transform as
~fi!V~fi; (20)
where V is an SU(2) matrix of the form V=ei~~b(~
f1;2;3g). So whileV0
ij(the singlet hopping and pairing)
is invariant under spin rotation, fV1
ij;V2
ij;V3
ijgtransforms as
a vector as expected since they represent triplet hopping and
pairing amplitudes.
B. PSG Classification
We now classify the non-Kramers spin-orbital liquids based
on projective representation similar to that of the conven-
tional quantum spin liquids.13Each spin-orbital liquid ground
state of the quadratic Hamiltonian (Eq 11) is character-
ized by the mean field parameters (eight on each bond,
;;E1;E2;E3;D2;D2;D3, or equivalently Uij). However,
due to the gauge redundancy of the spinon parametrization
(as shown in Eq. 15), a general mean-field ansatz need not
be invariant under the symmetry transformations on their own
but may be transformed to a gauge equivalent form without
breaking the symmetry. Therefore, we must consider its trans-
formation properties under a projective representation of the
symmetry group.13For this, we need to know the various pro-
jective representations of the lattice symmetries of the Hamil-
tonian (Eq. 4) in order to classify different spin-orbital liquid
states.
Operationally, we need to find different possible sets of
gauge transformations fGGgwhich act in combination withthe symmetry transformations fSGgsuch that the mean-field
ansatzUijis invariant under such a combined transformation.
In the case of spin rotation invariant spin-liquids (where only
the singlet channels andare present), the above statement
is equivalent to demanding the following invariance:
Uij= [GSS]Uij[GSS]y=GS(i)US(i)S(j)Gy
S(j);(21)
whereS2SGis a symmetry transformation and GS2GG
is the corresponding gauge transformation. The different pos-
siblefGSj8S2SGggive the possible algebraic PSGs that
can characterize the different spin-orbital liquid phases. To
obtain the different PSGs, we start with various lattice sym-
metries of the Hamiltonian. The action of various lattice
transformations15is given by
T1:(x;y;s )!(x+ 1;y;s);
T2:(x;y;s )!(x;y+ 1;s);
:(x;y;u )!(y;x;u );
(x;y;v )!(y;x;w );
(x;y;w )!(y;x;v );
S6:(x;y;u )!( y 1;x+y+ 1;v);
(x;y;v )!( y;x+y;w);
(x;y;w )!( y 1;x+y;u); (22)
where (x;y)denotes the lattice coordinates and s2fu;v;wg
denotes the sub-lattice index (see figure 2).
In terms of the symmetries of the Kagome lattice, these op-
erators obey the following conditions
T2=2= (S6)6=e;
g 1T 1gT=e8g2SG;
T 1
2T 1
1T2T1=e;
1T 1
1T2=e;
1T 1
2T1=e;
S 1
6T 1
2S6T1=e;
S 1
6T 1
2T1S6T2=e;
1S6S6=e: (23)
In addition, these commutation relations are valid in terms of
the operations on the pseudo-spin degrees of freedom, as can
be verified from Eq. 3.
In addition to the conditions in Eq. 23, the Hamiltonian is
trivially invariant under the identity transformation. The in-
variant gauge group (IGG) of an ansatz is defined as the set of
all pure gauge transformations GIsuch thatGI:Uij!Uij.
The nature of such pure gauge transformations immediately
dictates the nature of the low energy fluctuations about the
mean field state. If these fluctuations do not destabilize the
mean-field state, we get stable spin liquid phases whose low
energy properties are controlled by the IGG. Accordingly,
spin liquids obtained within projective classification are pri-
marily labelled by their IGGs and we have Z2;U(1)and
SU(2)spin liquids corresponding to IGGs of Z2;U(1)and5
SU(2)respectively. In this work we concentrate on the set of
Z2“spin liquids” (spin-orbital liquids with a Z2IGG).
We now focus on the PSG classification. As shown in Eq.
2, in the present case, the pseudo-spins transform non-trivially
under different lattice symmetry transformations. Due to the
presence of the triplet decoupling channels the non-Kramers
doublet transforms non-trivially under lattice symmetries (Eq.
3). Thus, the invariance condition on the Uijs is not given by
Eq. 21, but by a more general condition
Uij= [GSS]Uij[GSS]y=GS(i)S
US(i)S(j)
Gy
S(j):
(24)
Here
S
US(i)S(j)
=DSUS(i)S(j)Dy
S; (25)
andDSgenerates the pseudo-spin rotation associated with the
symmetry transformation ( S) on the doublet. The matrices
DShave the form
DS6= 1
20 ip
3
23; (26)
D=DT=i1;DT1=DT2= 0: (27)
Under these constraints, we must determine the relations
between the gauge transformation matrices GS(i)for our set
of ansatz. The additional spin transformation (Eq. 25) does
not affect the structure of the gauge transformations, as the
gauge and spin portions of our ansatz are naturally separate
(Eq. 12). In particular, we can choose to define our gauge
transformations such that
GS:Uij=GS:
ij !
ijGy
S(i) GS(j);
(28)
S:Uij=S:
ij !
S(i)S(j)DSDy
S ;(29)
where we have used the notation GS:UijGy
S(i)UijGS(j)
and so forth. As a result, we can build on the general con-
struction of Lu et al.15to derive the form of the gauge trans-
formation matrices. The details are given in Appendix B.
A major difference arises when examining the set of alge-
braic PSGs for Z2spin liquids found on the Kagome lattice
due to the difference between the structure of the time re-
versal symmetry operation on the Kramers and non-Kramers
pseudo-spin- 1=2s. In the present case, we find there are 30
invariant PSGs leading to thirty possible spin-orbital liquids.
This is in contrast with the Kramers case analysed by Lu et
al.,15where tenof the algebraic PSGs cannot be realized as
invariant PSGs, as all bonds in these ansatz are predicted to
vanish identically due to the form of the time reversal oper-
ator, and hence there are only twenty possible spin liquids.
However, with the inclusion of spin triplet terms and the non-
Kramers form of our time reversal operator, these ansatz are
now realizable as invariant PSGs as well. The time reversal
operator, as defined in Appendix B, acts as
T:
ij !~
ij ; (30)where ~=if2f1;2gand~= if2f0;3g.
The projective implementation of the time-reversal symmetry
condition (Eq. 23) takes the form (see Appendix B)
[GT(i)]2=TI8i; (31)
whereGT(i)is the gauge transformation associated with time
reversal operation and T=1for aZ2IGG.
Therefore, the terms allowed by the time reversal symmetry
to be non zero are, for T= 1,
10;11;12;13;20;21;22;23; (32)
and forT= 1, with the choice GT(i) =i 1(see appendix
B),
02;03;10;11;20;21;32;33: (33)
This contrasts with the case of Kramers doublets, in which no
terms are allowed for T= 1, and forT= 1the allowed
terms are
02;03;12;13;22;23;32;33: (34)
Further restrictions on the allowed terms on each link arise
from the form of the gauge transformations defined for the
symmetry transformations. All nearest neighbour bonds can
then be generated from Uijdefined on a single bond, by per-
forming appropriate symmetry operations.
Using the methods outlined in earlier works (Ref. 13, 15)
we find the minimum set of parameters required to stabilize Z2
spin-orbital liquids. We take into consideration up to second
neighbour hopping and pairing amplitudes (both singlet and
triplet channels). The results are listed in Table I.
The spin-orbital liquids listed from 21 30are not allowed
in the case of Kramers doublets and, as pointed out before,
their existence is solely due to the unusual action of the time-
reversal symmetry operator on the non-Kramers spins. Hence
these tenspin-orbital liquids are qualitatively new phases that
may appear in these systems. Of these ten phases, only two
(labelled as 21and22in Table I) require next nearest neigh-
bour amplitudes to obtain a Z2spin-orbital liquid. For the
other eight , nearest neighbour amplitudes are already suffi-
cient to stabilize a Z2spin-orbital liquid.
It is interesting to note (see below) that bond-pseudo-spin-
nematic order (Eq. 35 and Eq. 36) can signal spontaneous
time-reversal symmetry breaking. Generally, since the triplet
decouplings are present, the bond nematic order parameter for
the pseudo-spins21,22
Q
ij=h
S
iS
j+S
iS
j
=2 (~Si~Sj)=3i;(35)
as well as vector chirality order
~Jij=h~Si~Sji; (36)
are non zero. Since the underlying Hamiltonian Eq. 4) gener-
ally does not have pseudo-spin rotation symmetry, the above
non-zero expectation values do not spontaneously break any6
TABLE I: Symmetry allowed terms: We list the terms allowed to be non-zero by symmetry, for the 30 PSGs determined by Yuan-Ming Lu et
al15. The PSGs listed together are those with 12=1and all other factors equal. Included are terms allowed on nearest and next-nearest
neighbour bonds, as well as chemical potential terms which can be non zero on all sites for certain spin-orbital liquids. Also included is the
distance of bond up to which we must include in order to gap out the gauge fluctuations to Z2via the Anderson-Higgs mechanism13. Only
PSGs 9 and 10 can not host Z2spin-orbital liquids with up to second nearest neighbour bonds.
No. s n.n. n.n.n. Z2
1-2 2; 310;21;02;03;32;3310;21;02;03;32;33n.n.
3-4 010;21;02;03;32;3310;21n.n.
5-6 310;21;02;03;32;3310;21;03;33n.n.
7-8 0 11;2011;20;02;03;32;33n.n.n.
9-10 0 11;2011;20-
11-12 0 11;2010;11;02;32n.n.n.
13-14 310;11;03;3310;21;02;03;32;33n.n.
15-16 310;11;03;3310;21;03;33n.n.
17-18 010;11;03;3310;11;02;32n.n.
19-20 010;11;03;3310;21n.n.
21-22 010;21;22;2310;21;22;23n.n.n.
23-24 010;21;22;2310;11;12;23n.n.
25-26 011;12;13;2013;20;21;22n.n.
27-28 011;12;13;2010;11;13;22n.n.
29-30 011;12;13;2011;12;13;20n.n.
pseudo-spin rotation symmetry. However, because of the un-
usual transformation property of the non-Kramers pseudo-
spins under time reversal, the operators corresponding to
Q13
ij;Q23
ij;J1
ij;J2
ijare odd under time reversal, a symmetry of
the pseudo-spin Hamiltonian. Hence if any of the above op-
erators gain a non-zero expectation value in the ground state,
then the corresponding spin-orbital liquid breaks time rever-
sal symmetry. While this can occur in principle, we check
explicitly (see Appendix C) that in all the spin-orbital liquids
discussed above, the expectation values of these operators are
identically zero. This provides a non-trivial consistency check
on our PSG calculations.
We now briefly dicuss the effect of the fluctuations about
the mean-field states. In the absence of pairing channels (both
singlet and triplet) the gauge group is U(1). In this case, the
fluctuations of the gauge field about the mean field (Eq. 15)
are related to the scalar pseudo-spin chirality ~S1~S2~S3,
where the three sites form a triangle.19Such fluctuations are
gapless in a U(1)spin liquid. It is interesting to note that the
scalar spin-chirality is odd under time-reversal symmetry and
it has been proposed that such fluctuations can be detected
in neutron scattering experiments in presence of spin rotation
symmetry breaking.20In the present case, however, due to the
presence of spinon pairing, the gauge group is broken down
toZ2and the above gauge fluctuations are rendered gapped
through Anderson-Higg’s mechanism.13
In addition to the above gauge fluctuations, because of thetriplet decouplings which break pseudo-spin rotational sym-
metry, there are bond quadrupolar fluctuations of the pseudo-
spinsQ
ij(Eq. 35), as well as vector chirality fluctuations ~Jij
(Eq. 36)21,22on the bonds. These nematic and vector chirality
fluctuations are gapped because the underlying pseudo-spin
Hamiltonian (Eq. 4) breaks pseudo-spin-rotation symmetry.
However, we note that because of the unusual transformation
of the non-Kramers pseudo-spins under time reversal (only
thez component of pseudo-spins being odd under time re-
versal),Q13
ij;Q23
ij;J1
ijandJ2
ijare odd under time reversal.
Hence, while their mean field expectation values are zero (see
above), the fluctuations of these quantities can in principle lin-
early couple to the neutrons in addition to the z component
of the pseudo-spins.
Having identified the possible Z2spin-orbital liquids, we
can now study typical dynamic structure factors for these
spin-orbital liquids. In the next section we examine the typi-
cal spinon band structure for different spin-orbital liquids ob-
tained above and find their dynamic spin structure factor.7
FIG. 3: The spin structure factor for an ansatz in spin liquid 17, with
the spin variables transforming as a Kramers doublet.
FIG. 4: The spin structure factor for an ansatz in spin liquid 17, with
the spin variables transforming as a non-Kramers doublet.
IV . DYNAMIC SPIN STRUCTURE FACTOR
We compute the dynamic spin structure factor
S(q;!) =Zdt
2ei!tX
ijeiq(ri rj)X
a=1;2;3ha
i(t)a
j(0)i;
(37)
for an example ansatz of our spin liquid candidates, in order to
demonstrate the qualitative differences between the Kramers
and non-Kramers spin-orbital liquids. In the above equation,
the pseudo-spin variables are defined in a global basis (with
the z-axis perpendicular to the Kagome plane). In computing
the structure factor for the non-Kramers example, we include
only the3components of the pseudo-spin operator in the lo-
cal basis, since only the z-components carry magnetic dipole
moment (see discussion before). Hence, only this component
couples linearly to neutrons in a neutron scattering experi-
ment.
Eq. 37 fails to be periodic in the first Brillouin zone of
the Kagome lattice16, as the term ri rjin eq. 37 is a half-
integer multiple of the primitive lattice vectors when the sub-
lattices of sites i and j are not equal. As such, we examine the
structure factor in the extended brillouin zone, which consists
of those momenta of length up to double that of those in thefirst brillouin zone. We plot the structure factor along the cut
!M0!K0! , whereM0=2MandK0=2K. We
examine the structure factors for two ansatz of spin liquid #
17 which has both Kramers and non-Kramers analogues.
As expected, we find that the structure factor has greater
weight in the case of a Kramers spin liquid. This is par-
tially due to the fact that the moment of the scattering par-
ticle couples with all components of the spin, rather than sim-
ply thez-component. In addition, we note that the presence
of terms allowed in the non-Kramers spin-orbital liquid in-
duce the formation of a gap, which is absent for the Kramers
case with up to second nearest neighbour singlet and triplet
terms in this particular spin-orbital liquid. Qualitative and
quantitative differences such as these, which can be observed
in these structure factors between Kramers and non-Kramers
spin-orbital liquids, provides one possible distinguishing ex-
perimental signature of these states. We shall not pursue this
in detail in the present work.
V . DISCUSSION AND POSSIBLE EXPERIMENTAL
SIGNATURE OF NON-KRAMERS SPIN-ORBITAL LIQUIDS
In this work, we have outlined the possible Z2spin-orbital
liquids, with gapped or gapless fermionic spinons, that can be
obtained in a system of non-Kramers pseudo-spin-1/2s on a
Kagome lattice of Pr+3ions. We find a total of thirty , 10 more
than in the case of corresponding Kramers system, allowed
within PSG analysis in presence of time reversal symmetry.
The larger number of spin-orbital liquids is a result of the dif-
ference in the action of the time-reversal operator, when real-
ized projectively. We note that the spin-spin dynamic struc-
ture factor can bear important signatures of a non-Kramers
spin-orbital liquid when compared to their Kramers counter-
parts. Our analysis of the number of invariant PSGs leading to
possibly different spin-orbital liquids that may be realizable in
other lattice geometries will form interesting future directions.
We now briefly discuss an experiment that can play an im-
portant role in determining non-Kramers spin-orbital liquids.
Since the non-Kramers doublets are protected by crystalline
symmetries, lattice strains can linearly couple to the pseudo-
spins. As we discussed, the transverse ( xandy) components
of the pseudo-spins f1;2gcarry quadrupolar moments and
hence are even under the time reversal transformation. Fur-
ther, they transform under an Egirreducible representation
of the local D3dcrystal field. Hence any lattice strain which
has this symmetry can linearly couple to the above two trans-
verse components. It turns out that in the crystal type that
we are concerned, there is indeed such a mode related to the
distortion of the oxygen octahedra. Symmetry considerations
show that the linear coupling is of the form Eg11+Eg22
(fEg1;Eg2gbeing the two components of the distortion in
the local basis). The above mode is Raman active. For
a spin-liquid, we expect that as the temperature is lowered,
the spinons become more prominent as deconfined quasipar-
ticles. So the Raman active phonon can efficiently decay into
the spinons due to the above coupling channel. If the spin
liquid is gapless, then this will lead to anomalous broaden-8
ing of the above Raman mode as the temperature is lowered,
which, if observed, can be an experimental signature of the
non-Kramers spin-orbital liquid. The above coupling is for-
bidden in Kramers doublets by time-reversal symmetry and
hence no such anomalous broadening is expected.
Acknowledgments
We thank T. Dodds, SungBin Lee, A. Paramekanti and J.
Rau for insightful discussions. This research was supported
by the NSERC, CIFAR, and Centre for Quantum Materials at
the University of Toronto.
Appendix A: Crystal Field Effects
In this appendix, we explore the breaking of the J= 4spin
degeneracy by the crystalline electric field. The oxygen and
TM ions form a D3dlocal symmetry environment around the
Pr3+ions, splitting the ground state degeneracy of the elec-
trons. This symmetry group contains 6 classes of elements: E,
2C3,3C0
2,i,2S6, and 3d, where theC3are rotations by 2=3
about the local z axis, the C0
2are rotations by about axis per-
pendicular to the local z axis, iis inversion, S6is a rotation
by4=3combined with inversion and dis a reflection about
the plane connecting one corner and the opposing plane, run-
ning through the Prmolecule about which this is measured
(or, equivalently, a rotation about the x axis combined with
inversion). For our J=4 manifold, these have characters given
by
(4)(E) = 24 + 1 = 9 = (4)(i) (A1)
(4)(C3) =(4)(2
3) =sin(3)
sin(=3)= 0 =(4)(S6)(A2)
(4)(d) =(4)() =sin(9=2)
sin(=2)= 1 =(4)(C0
2)(A3)
where the latter equalities are given by the fact that our J=4
manifold is inversion symmetric. Thus, decomposing this in
terms ofD3dirreps, our l=4 manifold splits into a sum of
doublet and singlet manifolds as
l=4= 3Eg+ 2A1g+A2g: (A4)
To examine this further, we need to consider the matrix ele-
ments of the crystal field potential between the states of differ-
ent angular momenta. We know that this potential must be in-
variant under all group operations of D3d, so we can examine
the transformation properties of individual matrix elements,
hmjVjm0i. Under the C3operation, these states of fixed m
transform as
C3jmi=e2im
3jmi=!mjmi (!=e2i
3) (A5)
and thus the matrix elements transform as
C3:hmjVjm0i!hmj(C3) 1VC3jm0i=!m0 mhmjVjm0i:
(A6)By requiring that this matrix be invariant under this transfor-
mation, we can see that this potential only contains matrix
elements for mixing of states which have the z-component of
angular momentum which differ by 3. Thus, our eigenstates
are mixtures of the jm= 4i,jm= 1i, andjm= 2istates,
of thejm= 3i,jm= 0i, andjm= 3istates, and of the
jm= 4i,jm= 1i, andjm= 2istates.
In addition to this, we have the transformation properties
Tjmi= ( 1)mj mi (A7)
and
jmi= ( 1)mj mi (A8)
(where the operators for time reversal and reflection are
bolded for future clarity). Inversion acts trivially on these
states, as we have total angular momentum even. Thus our
time-reversal and lattice reflection (about one axis) symme-
tries give us doublet states of eigenstates jm= 4i+jm=
1i
jm= 2iandjm= 4i jm= 1i
jm= 2i
(with,,
2< in order to respect the time reversal symme-
try) for the three eigenstates of V in these sectors. The eigen-
states of thejm= 3i,jm= 0i, andjm= 3iportion of V
must therefore split into three singlet states, by our represen-
tation theory argument A4. Due to the expected strong Ising
term in our potential, we expect the eigenstate with maximal J
to be the ground state, meaning that to analyze the properties
of this ground state we are interested in a single doublet state,
one with large (close to one). We will restrict ourselves
to this manifold from this point forward, and define the two
states in this doublet as
j+i=jm= 4i+jm= 1i
jm= 2i (A9)
j i=jm= 4i jm= 1i
jm= 2i:(A10)
We shall also refer to states of angular momentum jm=ni
asjnifor simplicity of notation.
Appendix B: Gauge transformations
We begin by describing the action of time reversal on our
ansatz. The operation is antiunitary, and must be combined
with a spin transformation 1in the case of non-Kramers dou-
blets. As a result, the operation acts as T:
ij !
ij11 . However, we can simplify this consid-
erably by performing a gauge transformation in addition to
the above transformation, which yields the same transforma-
tion on any physical variables. The gauge transformation we
perform isi 2, which changes the form of the time reversal
operation to T:
ij !
ij11 2 2=
~
ij , where ~=if2f1;2gand~= if
2f0;3g.
On the Kagome lattice, the allowed form of the gauge trans-
formations has been determined by Yuan-Ming Lu et al.15For
completeness, we will reproduce that calculation, valid also9
for our spin triplet ansatz, here. The relations between the
gauge transformation matrices,
[GT(i)]2=TI; (B1)
G((i))G(i) =I; (B2)
Gy
T1(i)Gy
T(i)GT1(i)GT(T 1
1(i)) =T1TI; (B3)
Gy
T2(i)Gy
T(i)GT2(i)GT(T 1
2(i)) =T2TI; (B4)
Gy
(i)Gy
T(i)G(i)GT( 1(i)) =TI; (B5)
Gy
S6(i)Gy
T(i)GS6(i)GT(S 1
6(i)) =S6TI; (B6)
Gy
T2(T 1
1(i))Gy
T1(i)GT2(i)GT1(T 1
2(i)) =12I; (B7)
GS6(S 1
6(i))GS6(S 2
6(i))GS6(S3
6(i))
GS6(S2
6(i))GS6(S6(i))GS6(i) =S6I; (B8)
Gy
(T 1
2(i))Gy
T2(i)G(i)GT1((i)) =T1I; (B9)
Gy
(T 1
1(i))Gy
T1(i)G(i)GT2((i)) =T2I; (B10)
Gy
(S6(i))GS6(S6(i))G(i)GS6((i)) =S6I; (B11)
Gy
S6(T 1
2(i))Gy
T2(i)GS6(i)GT1(S 1
6(i)) =S6T1I;(B12)
Gy
S6(T 1
2T1(i))Gy
T2(T1(i))GT1(T1(i))
GS6(i)GT2(S 1
6(i)) =S6T2I; (B13)
are valid for our case as well, due to the decoupling of spin
and gauge portions of our ansatz. In the above, the relations
are valid for all lattice sites i= (x;y;s ), I is the 4x4 identity
matrix, and the GSmatrices are gauge transformation matri-
ces generated by exponentiation of the matrices. The ’s
are1, the choice of which characterize different spin liquid
states. In deriving this form of the commutation relations, we
have included a gauge transformation i 2in our definition of
the time reversal operator, as this simplifies the effect of the
operator on the mean field ansatz.
We turn next to the calculation of the gauge transforma-
tions. We look first at the gauge transformations associated
with the translations. We can perform a site dependent gauge
transformation W(i), under which the gauge transformations
associated with the translational symmetries transform as
GT1(i)!W(i)GT1(i)Wy(i ^x) (B14)
GT2(i)!W(i)GT2(i)Wy(i ^y): (B15)
As such, we can choose a gauge transformation W(i) to sim-
plify the form of GT1andGT2. Using such a transformation,
along with condition B7, we can restrict the form of these
gauge transformations to be
GT1(i) =iy
12I GT2(i) =I: (B16)
To preserve this choice, we can now only perform gauge
transformations which are equivalent on all lattice positions
(W(x;y;s ) =W(s)) or transformations which change the
shown matrices by an IGG transformation.
Next, we look at adding the reflection symmetry . Given
our formulae for GT1andGT2, along with the relations be-tween the gauge transformations, we have that
Gy
(T 1
2(i))G(i)x
12=T1I (B17)
Gy
(T 1
1(i))G(i)y
12=T2I: (B18)
DefiningG(0,0,s) =g(s), we have, by repeated application
of the above,
G(0;y;s) =y
T1g(s) (B19)
G(x;y;s ) =y
T1xy
12x
T2g(s): (B20)
Next, using
G((i))G(i) =I (B21)
we find that
I=G(y;x; (s))G(x;y;s ) (B22)
= (T1T2)x+yg((s))g(s): (B23)
Since this is true for all x and y, T1T2= 1 and thus
T1=T2andg((s))g(s) =I(where(u) =
u;(v) =wand(w) =v). Our final form for the gauge
transformation is
G(x;y;s ) =x+y
T1xy
12g(s): (B24)
Next we look at adding the S6symmetry to our calcula-
tion. We can do an IGG transformation, taking GT1(T1(i))
toS6T2GT1(T1(i)), with the net effect being that S6T2be-
comes one (previous calculations are unaffected). We now
have that
Gy
S6(T 1
2T1(i))GS6(i)y
12=I (B25)
Gy
S6(T 1
2(i))GS6(i) x 1
12 =S6T1I (s=u;v)(B26)
Gy
S6(T 1
2(i))GS6(i) x
12=S6T1I (s=w):(B27)
DefiningGS6(0;0;s) =gS6(s), we find that
GS6(n; n;s) =n(n 1)=2
12gS6(s) (B28)
GS6(x;y;s ) =x(x 1)=2+y+xy
12 x+y
S6T1gS6(s) (s=u;v)
(B29)
GS6(x;y;s ) =x(x 1)=2+xy
12 x+y
S6T1gS6(s) (s=w):
(B30)
Using the commutation relation between the andS6gauge
transformations, we find that
S6I=y
T1y
12y
S6T1gy
(v)gS6(v)g(u)gS6(u) (B31)
=y
T1y
12y
S6T1g(w)gS6(v)g(u)gS6(u)(B32)
giving us that T112S6T1 = 1 and
g(u)gS6(u)g(w)gS6(v) =S6I. A similar calcu-
lation on a different sublattice gives us
S6I=y
T1y
12y
S6T1gy
(w)gS6(w)g(v)gS6(w) (B33)
=y
T1y
12y
S6T1g(v)gS6(w)g(v)gS6(w)(B34)10
TABLE II: We list the solutions of Eq. B43 - B54, along with a set of gauge transformations which realize these solutions.
No.TTS6TS6S612g(u)g(v)g(w)gS6(u)gS6(v)gS6(w)
1,2 -1 1 1 1 1 11 0 0 0 0 0 0
3,4 -1 1 1 1 -1 11 0 0 0 0- 0i 1
5,6 -1 1 -1 1 -1 11 0 0 0i 3i 3i 3
7,8 -1 1 1 -1 -1 11i 1 0- 0 0i 1 0
9,10 -1 1 1 -1 1 11i 1 0- 0 0-i 1i 1
11,12 -1 1 -1 -1 1 11i 1 0- 0i 3-i 2i 3
13,14 -1 -1 -1 -1 -1 11i 3i 3i 3i 3i 3i 3
15,16 -1 -1 1 -1 1 11i 3i 3i 3 0 0 0
17,18 -1 -1 1 -1 1 11i 3i 3i 3 0 0i 1
19,20 -1 -1 -1 -1 1 11i 3i 3i 3i 3-i 3i 3
21,22 1 1 1 1 1 11 0 0 0 0 0 0
23,24 1 1 1 1 -1 11 0 0 0 0- 0i 3
25,26 1 1 1 -1 -1 11i 3 0- 0 0i 3 0
27,28 1 1 1 -1 1 11i 3 0- 0 0-i 3i 1
29,30 1 1 1 -1 1 11i 3 0- 0 0-i 3i 3
giving us (g(v)gS6(w))2=S6I. AZ2(IGG) gauge
transformation of the form W(x;y;s ) =y
T1changesT1
to 1. Using the cyclic relation of the gauge transformations
related to the S6operators, we find
S6I=12(gS6(w)gS6(v)gS6(u))2(B35)
giving us that
[gS6(w)gS6(v)gS6(u)]2=S612I: (B36)
Next we turn to the time reversal symmetry. Similar meth-
ods to the above give us that
[GT(i)]2=TI (B37)
Gy
T(i)GT(i+ ^x) =T1TI (B38)
Gy
T(i)GT(i+ ^y) =T2TI: (B39)
The first of these relations tells us that GT(i)is either the
identity (for T= 1) ori~ a~ (forT= 1, wherej~ aj= 1.
DefiningGT(0;0;s) =gT(s),
GT(x;y;s ) =x
T1Ty
T2TgT(s) (B40)
and further, using the commutation relations between the
andTgauge transformations and the S6andTgauge trans-
formations,
gy
(s)gy
T(s)g(s)gT((s))x+y
T1Tx+y
T2T=TI(B41)
gy
S6(s)gy
T(s)gS6(s)gT(S 1
6(s))f1(i)
T1Tf2(i)
T2T=S6TI:(B42)Because this is true for all x and y, and f1(i)is not equal to
f2(i),T1T=T2T= 1. IfGT(i) =i~ a~ , we perform a
gauge transformation W on GT(i)such thatWyGT(i)W=
i1(as this is the same on all sites, it does not affect our gauge
fixing for the translation gauge transformations). Collecting
the necessary results for further use,
GT1(x;y;s ) =y
12I (B43)
GT2(x;y;s ) =I (B44)
G(x;y;s ) =xy
12g(s) (B45)
GS6(x;y;s ) =xy+(x+1)x=2
12 gS6(s)s=u;v (B46)
GS6(x;y;s ) =xy+x+y+(x+1)x=2
12 gS6(s)s=w(B47)
GT(s) =I=gT(s)T= 1 (B48)
GT(s) =i 1=gT(s)T= 1 (B49)
g((s))g(s) =I (B50)
g(u)gS6(u)g(w)gS6(v) = (g(v)gS6(w))2=S6I
(B51)
(gS6(w)gS6(v)gS6(u))2=S612I (B52)
g(s)gT((s)) =TgT(s)g(s) (B53)
gS6(s)gT(S 1
6(s)) =S6TgT(s)gS6(s): (B54)
We also have the gauge freedom left to perform a gauge ro-
tation arbitrarily at all positions for T= 1 or an arbitrary
gauge rotation about the x axis for T= 1.
The solution to the above equations is derived in detail by
Luet al.15and as such we simply list the results in table II.
The basic method of obtaining these solutions is as follows:11
for each choice of Z2parameter set, we determine whether
there is a choice of gauge matrices fgSgwhich satisfy the
equations B43 - B54. In order to do so, we determine the
allowed forms of the gSmatrices from the equations, then use
the gauge freedom on each site to fix the form of these. Of
particular not is the fact that in the consistency equations for
thegmatrices, the terms 12andS6only appear multiplied
together, meaning that for any choice of the gauge matrices
gSwe can choose 12=1, which fixes the form of S6.
Appendix C: Relation among the mean-field paramters
The relation among the different singlet and triplet param-
eters in terms of ijis given by
ij=00
ij+03
ij;ij= 01
ij+i02
ij;
E1
ij=10
ij+13
ij;E2
ij=20
ij+23
ij;E3
ij=30
ij+33
ij
D1
ij= 11
ij+i12
ij;D2
ij= 21
ij+i22
ij;D3
ij= 31
ij+i32
ij
(C1)
Using these, we can derive the form of the bond nematic
order parameter and vector chirality order parameters, which
are given in terms of the mean field parameters21as
Q;
ij= 1
2
E
ijE
ij 1
3;j~Eijj2
+h:c:
1
2
D
ijD
ij 1
3;j~Dijj2
+h:c:
J
ij=i
2
ijE
ij
ijE
ij
+i
2
ijD
ij
ijD
ij
(C2)where our definition of ijdiffers by a factor of (-1) from that
of the cited work. We rewrite this in terms of our variables,
finding
Q
ij= 0
ij0
ij+X
aa
ija
ij
+
3X
b
(b0
ij)2 X
a(ba
ij)2)
J
ij=i(00
ij0
ij X
a0a
ija
ij) (C3)
In particular, we find that J1,J2,Q13andQ23must be zero
for all non-Kramers spin liquids, as the terms allowed by sym-
metry in Eq. 32 and 33 do not allow non-zero values for these
order parameters.
1P. Fazekas, Lecture Notes on Electron Correlation and Magnetism
(Series in Modern Condensed Matter Physics) , World Scientific
Pub Co Inc (1999).
2K. Yosida, Theory of Magnetism , Springer (2001).
3B. Bleaney and H. E. D. Scovil, Phil. Mag. 43, 999 (1952).
4Shigeki Onoda and Yoichi Tanaka, Phys. Rev. B 83, 094411(R)
(2011).
5K. Matsuhira, Y . Hinatsu, K. Tenya1, H. Amitsuka and T. Sakak-
ibara, J. Phys. Soc. Jpn. 71, 1576 (2002).
6K. Matsuhira, C. Sekine, C. Paulsen, M. Wakeshima, Y . Hinatsu,
T. Kitazawa, Y . Kiuchi, Z. Hiroi, and S. Takagi, J. Phys. Conf. Ser.
145, 012031 (2009).
7S. Nakatsuji, Y . Machida, Y . Maeno, T. Tayama, T. Sakakibara, J.
van Duijn, L. Balicas, J. N. Millican, R. T. Macaluso, and J. Y .
Chan, Phys. Rev. Lett. 96, 087204 (2006).
8SB. Lee, S. Onoda, and L. Balents, Phys. Rev. B 86, 104412
(2012).
9R. Flint, and T. Senthil, arXiv:1301.0815 (2013).
10P. Chandra, P. Coleman, and R. Flint, Nature 493, 621 (2013).
11O. Suzuki, H. S. Suzuki, H. Kitazawa, G. Kido, T. Ueno, T. Ya-maguchi, Y . Nemoto, T. Goto, J. Phys. Soc. Japan 75, 013704
(2006).
12H. A. Kramers, Proc. Amsterdam Acad. 33, 959 (1930).
13X. G. Wen, Phys. Rev. B 65, 165113 (2002); X.-G. Wen, Quantum
Field Theory of Many-Body Systems. Oxford University Press,
Oxford (2004).
14P. W. Anderson, G. Baskaran, Z. Zou, and T. Hsu, Phys. Rev. Lett.
58, 2790 (1987).
15Yuan-Ming Lu, Ying Ran, and Patrick A. Lee, Phys. Rev. B 83,
224412 (2011).
16T. Dodds, S. Bhattacharjee, and Y . B. Kim, arXiv:1303.1154
17R. Schaffer, S. Bhattacharjee, and Y . B. Kim, Phys. Rev. B 86,
224417 (2012).
18I. Affleck, and J. B. Marston, Phys. Rev. B 37, 3774 (1988).
19P. A. Lee, and N. Nagaosa, Phys. Rev. B 46, 5621 (1992).
20P. A. Lee, and N. Nagaosa, Phys. Rev. B 87, 064423 (2013).
21R. Shindou, and T. Momoi, Phys. Rev. B 80, 064410 (2009).
22S. Bhattacharjee, Y . B. Kim, S.-S. Lee, and D.-H. Lee, Phys. Rev.
B 85, 224428 (2012). |
1609.03141v1.Spin_Orbit_Coupling_Induced_Spin_Squeezing_in_Three_Component_Bose_Gases.pdf | Spin-Orbit Coupling Induced Spin Squeezing in Three-Component Bose Gases
X. Y. Huang,1F. X. Sun,1W. Zhang,2, 3,Q. Y. He,1, 4,yand C. P. Sun5,z
1State Key Laboratory of Mesoscopic Physics, School of Physics, Peking University,
Collaborative Innovation Center of Quantum Matter, Beijing 100871, China
2Department of Physics, Renmin University of China, Beijing 100872, China
3Beijing Key Laboratory of Opto-electronic Functional Materials and Micro-nano Devices,
Renmin University of China, Beijing 100872, China
4Collaborative Innovation Center of Extreme Optics,
Shanxi University, Taiyuan, Shanxi 030006, China
5Beijing Computational Science Research Center, Beijing 100084, China
Weobservespinsqueezinginthree-componentBosegaseswhereallthreehyperfinestatesarecoupled
by synthetic spin-orbit coupling. This phenomenon is a direct consequence of spin-orbit coupling, as
can be seen clearly from an effective spin Hamiltonian. By solving this effective model analytically
with the aid of a Holstein-Primakoff transformation for spin-1 system in the low excitation limit,
we conclude that the spin-nematic squeezing, a novel category of spin squeezing existing exclusively
in large spin systems, is enhanced with increasing spin-orbit intensity and effective Zeeman field,
which correspond to Rabi frequency
Rand two-photon detuning within the Raman scheme
for synthetic spin-orbit coupling, respectively. These trends of dependence are in clear contrast
to spin-orbit coupling induced spin squeezing in spin-1/2 systems. We also analyze the effects of
harmonic trap and interaction with realistic experimental parameters numerically, and find that a
strong harmonic trap favors spin-nematic squeezing. We further show spin-nematic squeezing can be
interpreted as two-mode entanglement or two-spin squeezing at low excitation. Our findings can be
observed in87Rb gases with existing techniques of synthetic spin-orbit coupling and spin-selectively
imaging.
I. INTRODUCTION
Spin squeezing is an important resource which has
many potential applications not only in quantum metrol-
ogy and atom interferometers [1–5], but also in many
aspects of quantum information due to its close rela-
tion with quantum entanglement [6–10]. In conventional
experiments, squeezing is usually achieved via the non-
linearity induced by the inter-particle interaction [3–5].
As an example, spin squeezing has been obtained ex-
perimentally in a Bose-Einstein condensate (BEC) of a
three-component Bose gas [11]. However, the intensity
of spin squeezing in these experiments crucially depends
on the interaction between atoms. In cold atom exper-
iments, the background interaction is usually very weak
such that the observation of squeezing is relatively hard.
Although there are some techniques to enhance the in-
teraction, e.g., by tuning the state-dependent microwave
potentials [4], or through a magnetic Feshbach resonance
in alkali atoms [5, 12], the side effects of decoherence,
severe atom loss and dynamical instability induced by
strong interaction still hinder the achievement of strong
spin squeezing.
The experimental realization of synthetic spin-orbit
coupling (SOC) in ultracold atomic gases [13–15] has at-
tracted much attention, partly due to its close relation
wzhangl@ruc.edu.cn
yqiongyihe@pku.edu.cn
zcpsun@csrc.ac.cnto exotic many-particle states and novel excitations [16–
18]. Recently, theoreticalstudieshaveproposedtorealize
spin squeezing in two-component BEC by synthetic spin-
orbit coupling (SOC) [19, 20]. It has been shown that the
presence of SOC will induce an effective spin-spin inter-
action which can lead to spin squeezing. However, there
are two disadvantages of these proposals. First, the syn-
thetic SOC requires a Raman transition between two hy-
perfine states. The Rabi frequency of this Raman transi-
tion is detrimental to spin squeezing, i.e., a stronger SOC
leads to a weaker squeezing. Besides, the two-photon de-
tuning of this Raman transition is also unfavorable such
that best squeezing will be achieved when the detuning is
zero. Nonetheless, in realistic experiments one would en-
counter severe heating effect when the detuning is tuned
on resonance.
In this paper, we study spin squeezing in a three-
component Bose gases where all three hyperfine states
are coupled by spin-orbit coupling induced by Raman
transitions. As a result, this system has pseudo-spin-1,
andthespinoperatorshereinmustbedescribedbySU(3)
spin matrices, i.e., the Gell-Mann matrices. These Gell-
Mann matrices span an eight-dimensional spin hyper-
space, with three of them are usually refereed as spin
vectors, and the other five as nematic tensors [21]. The
squeezed spin operators hence can be categorized into
three types, including the spin-spin squeezing, nematic-
nematic squeezing, and spin-nematic squeezing. Here,
we focus on the spin-nematic squeezing, as it is a novel
typeofsqueezingwhichexistsexclusivelyinsystemswith
large spins. We find that the presence of SOC can induce
spin-nematic squeezing, which can be further enhancedarXiv:1609.03141v1 [quant-ph] 11 Sep 20162
byincreasingtheSOCintensityorreducingthequadratic
Zeeman splitting. These trends of dependence can be
understood from an effective Hamiltonian, in which the
Rabi frequency and quadratic Zeeman splitting corre-
spond to effective Zeeman fields in the spin and nematic
sectors, respectively, hence causing opposite effects on
various types of spin squeezing. More importantly, we
find that the squeezing is favored by two-photon detun-
ing of the Raman transition within a fairly large param-
eter regime, which is beneficial for experimental realiza-
tions to avoid severe heating effect. When the system ex-
hibits spin-nematic squeezing in the low excitation limit,
we also find two-mode entanglement [22] and two-spin
squeezing [23]inthesystem. Wefurtherstudytheeffects
of an external trapping potential and inter-atomic inter-
action which are present in realistic experimental situa-
tionsbynumericallyanalysis, andconcludethatthespin-
nematic squeezing is favored by stronger trapping poten-
tials. Finally, we discuss possible detection scheme via
a spin-selective imaging technique and a radio-frequency
(RF) rotation of the spin axes [11].
The remainder of this paper is organized as follows. In
Sec. II, we introduce the system under investigation and
discuss the single-particle spectra. We then derive an ef-
fectivespinHamiltonianfromwhichitcanbeseenclearly
that SOC induces an effective spin-spin interaction. We
then analyze the spin-nematic squeezing and its depen-
dence of various factors in Sec. III. Finally, we discuss
possible experimental detection scheme and summarize
in Sec. IV.
II. SINGLE-PARTICLE SPECTRA AND
EFFECTIVE HAMILTONIAN
Spin-orbit coupled three-component Bose gas can be
generalized by counter-propagating Raman lasers along
^xto couple the three hyperfine states with momentum
transfer of the Raman process 2kr. The non-interacting
Hamiltonian can be written in the matrix form as [24]
H=0
BBBBB@(kx+2kr)2
2
R=2 0
R=2k2
x
2
R=2
0
R=2(kx 2kr)2
2+1
CCCCCA+k2
?
2;
(1)
wherek?=q
k2y+k2zis the transverse momentum, is
the two-photon detuning from the Raman resonance, is
the quadratic Zeeman shift induced by the magnetic field
along ^y, and
RrepresentstheRabifrequencyoftheRa-
man transition. Notice that throughout the manuscript,
we use the natural units of }=m= 1, and define krand
the recoil energy Er=k2
r=2as the units of momentum
and energy, respectively.
The single-particle dispersion can be obtained by di-
agonalizing th non-interacting Hamiltonian of Eq. (1).
1minimum
1minimum2minima
2minima3minimaæcritical point
ÑdEr=1HaL
0 1 2 3 4 5 6-6-4-20246
ÑWRErÑeEr
1minimum1minimum
1minimum2minima
2minimaæ
æcritical point
critical point
ÑeEr=6HbL
0 2 4 6 8-10-50510
ÑWRErÑdEr
HcL
ÑWREr=76543210
-4-2024-4-3-2-1012
kxkrEEr
ÑeEr=6543210-1-2-3HdL
-4-2024-6-4-202
kxkrEErFigure 1. (Color online) (a-b) Single-particle phase diagrams
of a three-component Bose gas with one-dimensional SOC in
the (a)
R–plane with = 1and (b)
R–plane with
= 6. The lowest branch of the single-particle dispersion
spectrum acquires either one, two, or three local minima in
different parameter regimes separated by solid lines. On the
dashed lines within regions of multiple minima, two of the
local minima are degenerate. Typical examples for the lowest
branch of dispersion curves by changing (c) Rabi frequency
Rwith= 1and= 0and (d) quadratic Zeeman energy
with= 1and
R= 2.
The resulting spectra has three branches, among which
thelowestonecanhavethreeminima,twominima,orone
minimum depending on the combination of parameters.
In Figs. 1(a) and 1(b), we show the parameter regions ex-
hibitingdifferentstructuresforthecaseof ~=Er= 1and
~=Er= 6respectively. From Fig. 1(a), we can identify
various regions where the lowest branch of single-particle
dispersion acquires 1, 2, or 3 minima. Specifically, for
the case of a large positive quadratic Zeeman splitting
, thej0istate is far detuned from the other two high-
lying hyperfine states, so that the spectrum has only one
minimum. On the other hand, if is large negative, the
j0istate becomes the high-lying state and the system
essentially turn into a spin-1/2 Bose gas where the two
j1ispin components are spin-orbit coupled via virtual
processes involving the j0istate. As a result, the single-
particle dispersion can have either two or one minima,
depending on the SOC intensity
Rand two-photon de-
tuning. For the case of intermediate jj, all three hyper-
fine states are spin-orbit coupled and the shape of spec-
trum is sensitively dependent on all parameters. Typical
examples of dispersion spectra along the kxaxis showing
one minimum, two minima, and three minima, as well
as the trends of evolution depending on
Randare
illustrated in Figs. 1(c) and 1(d), respectively.
As the spin operators in spin-1/2 systems all belong
to the SU(2) group, those in spin-1 systems discussed3
here are elements in the SU(3) group. The SU(3) group
is locally isomorphic to the O(8) group, which has eight
linearly independent observables as generators. These
generatorscanbegroupedintotwotypes, includingthree
spin vectors (or angular momentum operators) and five
nematic tensors. The irreducible matrix representations
of these observables are given by [21]
Jx=1p
20
@0 1 0
1 0 1
0 1 01
A,Jy=ip
20
@0 1 0
1 0 1
0 1 01
A,
Jz=0
@1 0 0
0 0 0
0 0 11
A,Qxy=i0
@0 0 1
0 0 0
1 0 01
A,
Qyz=ip
20
@0 1 0
1 0 1
0 1 01
A,Qzx=1p
20
@0 1 0
1 0 1
0 1 01
A,
D=0
@0 0 1
0 0 0
1 0 01
A,Y=1p
30
@1 0 0
0 2 0
0 0 11
A:
The commutators between these spin operators can then
be classified into three categories: [Jy;Jz] =iJxas
spin-spin group, [Qxy;Qxz] =iJx,[Qyz;D] =iJx,
and [Qyz;Y] =p
3iJxas nematic-nematic group, and
[Jx;Qyz] =i(p
3Y+D),[Jy;Qzx] =i( p
3Y+D)as
spin-nematic group.
To study the effective spin-spin interaction induced by
SOC,aswellastheinducedspinsqueezingeffect, nextwe
derive an effective spin model. To facilitate the deriva-
tion, we impose a weak harmonic trap V(x) =!2
xx2=2 +
!2
yy2=2 +!2
zz2=2. We will find that the resulting form
of the effective model does not depend on the absolute
value of trapping frequency, hence incorporate solely the
effect of SOC. In the presence of such an auxiliary trap-
ping potential, we can quantize the motional degrees of
freedom along the trapping direction to a discrete energy
spectrum. In particular, by introducing the bosonic op-
eratorsap
!x=2(x+ikx=!x),bp
!y=2(y+iky=!y),
cp
!z=2(z+ikz=!z)and the collective spin operators
Fs=x;y;z =PN
i=1Ji;s,FY=PN
i=1Yi, the Hamiltonian of
Eq. (1) for N-particle can be rewritten as
~H=!xNaya+N4k2
r
3+
Rp
2Fx
+ikrp
2!x(ay a)Fz Fz+2k2
r+p
3FY:(2)
Here we ignore !yNbyb+!zNcycsince the boson modes
iny,zdirection do not interact with the ultracold atoms.
Employing the unitary transformation U= exp[iG(ay+a)Fz]withG=p
2=!xkr=N, the Hamiltonian thus can
be transformed as
~H0=!xNaya qF2
z Fz+2k2
r+p
3FY
+
Rp
2fFxcos[G(ay+a)] Fysin[G(ay+a)]g;(3)
whereq= 4k2
r=N= 8Er=N. Notice that the term of
N(4k2
r )=3has been dropped out as the zero-point
energy.
For a BEC, the expectation value of hayaiis in the
order ofNfor the ground state, and about unity for
excited states. Considering the prefactor of 1=Nin the
definition of G, the leading order of the arguments in the
cosine and sine functions in Eq. (3) are 1=p
N, which
is negligible for systems of large particle number. As a
result, we can approximate the cosine and sine functions
tothezerothorder,andtheHamiltonianEq. (3)becomes
separable in spatial and spin degrees of freedom, leading
to an effective spin Hamiltonian
He= qF2
z+
Rp
2Fx Fz+4Er+p
3FY:(4)
One can see clearly that an effective spin-spin interac-
tion emerges as a result of SOC, and the Rabi frequency
R, two-photon detuning , and the quadratic Zeeman
splittingact as effective Zeeman fields along different
directions in the eight-dimensional spin hyperspace.
III. SPIN-NEMATIC SQUEEZING
With the aid of the effective spin model of Eq. (4), we
can study the spin squeezing in the underlying system.
As the commutators relation between spin and nematic
operators are not present in the spin-1/2 case, next we
focus on spin squeezing of this type. The method can be
straightforwardly applied to the spin-spin and nematic-
nematic commutators, and the results are qualitatively
consistent with the findings for the spin-spin case in spin-
1/2 system with SOC [19, 20].
The spin model of Eq. (4) can not be solved analyt-
ically due to the presence of nonlinear interaction. In
the low excitation limit, however, we can introduce the
Holstein-Primakoff transformation for spin-1 systems
Fx1p
2
by
1N0
0+N0
0b 1+ h:c:
;
Fy1p
2i
by
1N0
0+N0
0b 1 h:c:
; (5)
whereN0
0q
N by
1b1 by
1b 1, and the operators b1
andb 1representing spin flipping processes between the
internal levelsj1iandj0i, represented by the bosonic
modesa1anda0. For the case that most of the par-
ticles remain in the mode a0, i.e.,hay
0a0i 'Nand4
hby
1b1iN, the operators b1=a1ay
0=p
Nare effec-
tive bosonic modes satisfying the bosonic commutation
relationsh
b;by
i
=with; =1. Within the
assumption that the majority of the particles are resid-
ing in thej0istate, or equivalently the excitations to
thej1istates are rare, we can rewrite the bosonic
operators as a mean-field value plus some fluctuations
b1=p
N1+b1. The ground state energy can
the be obtained by minimizing the energy functional
E(1; 1). As in this the low excitation limit, nearly
all the spins are polarized in FYdirection, which means
p
3FY+FD2N, the squeezing parameter is then
given by [25]
xmin
42Jn?
J=242Fx=N; (6)
Here,Jis the expectation value of mean spin, Jn?is a
spin component along the direction perpendicular to the
mean spin direction. So in our case, it is clear that x
can be obtained by calculated the variance of Fx, one has
spin squeezing in the spin-nematic channel as x<1.
We first discuss the case of zero two-photon detuning
= 0, and show in Fig. 2 the spin-nematic squeezing pa-
rameter as functions of Rabi frequency
Rand quadratic
Zeeman splitting . One can see clearly that the ground
stateis aspinsqueezed stateundertheeffect ofSOC.Im-
portantly, as shown in Fig. 2(a), spin-nematic squeezing
can be enhanced with increasing
R. This behavior is in
stark contrast to the case of spin-1/2 systems, where the
spin-spin squeezing is favored by decreasing
R[19, 20].
We then extend the discussion to the more general case
of a nonzero two-photon detuning 6= 0. This scenario is
experimentally relevant because a severe heating effect is
usually present as the Raman transition is on-resonance.
As shown in Fig. 3(a), a finite favors spin-nematic
squeezingwithinafairlylargeregionof j~=Erj<5. This
result can be understood by analyzing the single-particle
Hamiltonian of Eq. 1, where andare energy offsets
of the diagonal elements. As Raman transitions will be
enhanced when difference states are near resonance, spin
squeezing will be favored when the absolute value of
is close to. To further clarify this argument, we ana-
lyze the atom populations of different ground states with
changing. As shown in Fig. 3(b), the presence of a fi-
nitewill enhance the transition between the j0istate
and one of thej1istates, while the transition to the
otherj1istate is reduced. Notice that this behavior
is very different from the spin-1/2 case, where the two
spin components are moved away from each other with
increasing, leading to an effectively weaker SOC.
The dependences of spin-nematic squeezing on the var-
ious parameters of
R,andcan also be interpreted
from the effective spin model of Eq. (4), within which the
three parameters correspond to effective Zeeman fields
along theFx,FY, andFydirections, respectively. Con-
sidering that in the low excitation limit nearly all spins
are polarized along the FYdirection, a stronger Zeeman
HaL
012340.70.80.91.0
ÑWRErxx
HbL
2468100.70.80.91.0
ÑeErxxFigure 2. (Color online) Spin-nematic squeezing parameter x
as a function of (a) Rabi frequency
Rwith= 0and= 6
and (b) quadratic Zeeman splitting with= 0and
R= 2.
In both figures, results obtained from the effective spin model
Eq.(4) are illustrated by blue solid lines, in comparison to the
numerical solutions of the GP equation for a pancake-shaped
trapwith!x=!y= 50Hz,!z= 1500Hz(blackdashed), and
for a cigar-shaped trap with !x=!y= 5000Hz,!z= 1500
Hz (red dotted). Here, we consider a gas of87Rb atoms in
theF= 1manifold with background interaction and total
particle number N= 105.
field along the same direction, i.e., a larger value of ,
will further intensify the polarization so that the effec-
tive spin-spin interaction becomes relatively weak, lead-
ing to a less spin-nematic squeezing effect. On the other
hand, effective Zeeman fields along the perpendicular di-
rections, either FxorFz, will tilt the spin polarization
from theFYaxis slightly but the effect spin-spin interac-
tion is enhanced obviously, resulting an increased squeez-
ing parameter as in Eq. (6).
In addition to the spin-nematic squeezing, we notice
that in the low excitation limit with the majority of par-
ticles residing in the j0istate, the two effective bosonic
modesb1andb 1can be entangled, which is referred
as two-mode entanglement. A sufficient criterion for en-
tanglement between the modes b1andb 1from the spin
squeezing parameters is then given by [22]
DCZ= (
+++=2
)=2<1; (7)
where
h2F
i=Nrepresentsthevarianceofquadra-
ture phase amplitudes which depends on the parameter5
HaL
-5.0-2.50.02.55.00.750.800.850.900.951.00
ÑdErxx
HbL
-10-505100.000.200.400.600.801.00
ÑdErr
Figure 3. (Color online)(a) Variations of spin-nematic squeez-
ing parameter as a function of two-photon detuning with
R= 2and= 6. Analytic result obtained from the effective
spin model Eq. (4) within low-density excitation approxima-
tion (blue solid) is compared with numerical solutions of the
GP equation for a pancake-shaped trap with !x=!y= 50
Hz,!z= 1500Hz (black dashed), and for a cigar-shaped trap
with!x=!y= 5000Hz,!z= 1500Hz (red dotted). (b)
Atom number fractions of the j 1i(black dotted),j0i(blue
dashed), andj+ 1i(red solid) states.
, andwehaveusethedefinitions F
+= cosFx+sinFyz
andF
= cosFzx+ sinFyin this system. Here, the
collective nematic operators are Fyz=PN
i=1Qyzand
Fzx=PN
i=1Qzx. Figure 4(a) shows that
DCZreaches
its minimum for =nwithnan integer, and the en-
tanglement is enhanced by Raman transition.
Another representation of two-mode entanglement is
called two-spin squeezing, which is defined by dividing
the spin-1 space into three subspaces pseudospins (each
of spin-1/2) U,VandTassociated the three relative
number differences of particles N+1 N0,N 1 N0, and
N+1 N 1in three-component labeled by f+1; 1;0g
[23]. Two-spin squeezing parameter is given to describe
the correlation between the spin subspace U(the spin
flipping process between internal levels j+ 1iandj0i)
andV(the spin flipping process between internal levels
j 1iandj0i) [23]
UV=2F
++ 2F+=2
p
3jhFYij<1; (8)
In Fig. 4(b), one can see clearly that the optimal correla-
0p
2p3p
22p0.920.961.00
qxDCZqHaL
0 1 2 3 40.700.800.901.00
ÑWRErxDCZ0
0p
2p3p
22p0.961.001.04
qxUVqHbL
0 1 2 3 40.850.900.951.00
ÑWRErxuv0Figure 4. (Color online) (a) Two-mode entanglement param-
eter0
DCZand (b) two-spin squeezing parameter 0
UVver-
sus Rabi frequency
Rwith other parameters being = 0
and= 6. Analytic result obtained from the effective spin
model Eq. (4) within low-density excitation approximation
(blue solid) is compared with numerical solutions of the GP
equation for a pancake-shaped trap with !x=!y= 50Hz,
!z= 1500Hz (black dashed), and for a cigar-shaped trap
with!x=!y= 5000Hz,!z= 1500Hz (red dotted). The
insets show the squeezing parameters as functions of . No-
tice that the optimal squeezing in both criteria are obtained
when=nwithnan integer.
tion is obtained when =nwithnan integer, and in-
creases with the the Raman transition. When comparing
spin-nematic squeezing parameter (Fig. 2(a)) with these
two criterions (Fig. 4), we find that the effect of squeez-
ing in spin-nematic channel is another representation of
the correlation between two spin subspaces and entangle-
ment between two effective modes in the low excitation
limit.
Finally, we notice that in realistic experiments, one
also needs to take the effects of inter-atomic interaction
and a global harmonic trap into consideration. Taking
87Rb as a particular example, the interaction among the
three hyperfine states of the ground state manifold can
be categorized into two groups, depending on the total
angularmomentumofthetwocollidingatoms. Theback-
ground scattering lengths are taken as as0= 101:8a0for
F= 0, andas2= 100:4a0forF= 2, wherea0denotes the
Bohr radius [26]. For the effects of trapping potentials,
we consider two types of global harmonic traps includ-
ing a pancake-shaped quasi-two-dimensional trap with6
!x=!y= 50Hz and!z= 1500Hz, and a cigar-shaped
three-dimensional trap with !x=!y= 5000 Hz and
!z= 1500Hz.
BynumericallysolvingtheGross-Pitaevski(GP)equa-
tion for a total number of N= 105atoms, we obtain
the ground state of the system, and calculate the spin-
nematic squeezing parameter x, the two-mode entangle-
ment parameter
DCA, and the two-spin squeezing pa-
rameter
UV. The corresponding results are shown in
Figs.2, 3and4. Bycomparingthenumericalresultswith
the outcome from the effective spin model, we conclude
that the effective model Eq. (4) is qualitatively valid in
thelowexcitationlimit. Ontheotherhand, astronghar-
monic trap can cause sizable increment on spin-nematic
squeezing and two-mode entanglement. This observation
can be understood by noticing that in the presence of
a strong harmonic trap, the particles will be more con-
densed with a higher number density at the trap center.
As a result, the inter-particle interaction has stronger
effect and causes better spin-nematic squeezing and two-
mode entanglement.
IV. EXPERIMENTAL DETECTION AND
CONCLUSION
We have shown that an effective spin-spin interac-
tion can be induced in spin-orbit coupled spin-1 BEC,
whichcanproduceaspecialkindofsqueezingcalledspin-
nematicsqueezing. Thistypeofspinsqueezingcanbeen-
hanced by increasing Raman transition intensity and de-creasing quadratic Zeeman splitting. More importantly,
the squeezing is favored by a finite two-photon detuning
in a fairly large parameter regime, which could be bene-
ficial for experiments to reduce heating effect. These be-
haviors are in clear contrast to the spin squeezing within
spin-orbit coupled spin-1/2 systems, where the trends
of dependence on Raman transition intensity and two-
photon detuning are opposite. We also observe SOC
induced two-mode entanglement and two-spin squeezing
in such a system, and investigate their dependence on
Raman transition intensity. We further analyze the ef-
fects of inter-particle interaction and external harmonic
trap by numerically solving the GP equation, and find
good agreement with approximate solutions of the effec-
tive spin model.
In order to detect such an exotic type of spin squeezing
in this system, one may need to rotate Jxinto the easily
measuredJzdirection by applying a =2radio-frequency
(RF) rotation about the Jyaxis. This operation can be
accomplished with a two-turn coil on the experimental
y-axis driven at the frequency splitting of the mFstates.
Then, we can measure the variance of spin via a spin-
selective imaging technique.
ACKNOWLEDGMENTS
This work is supported by NSFC (11274009, 11622428,
11274025,11434011,11522436,61475006,and61675007),
NKBRP (2013CB922000) and the Research Funds of
Renmin University of China (10XNL016, 16XNLQ03).
[1] D. J. Wineland, J. J. Bollinger, W. M. Itano, and D. J.
Heinzen, Phys. Rev. A 50, 67 (1994).
[2] M.KitagawaandM.Ueda, Phys.Rev.A 47, 5138(1993).
[3] J. Estève, C. Gross, A. Weller, S. Giovanazzi, and M. K.
Oberthaler, Nature 455, 1216 (2008).
[4] M. F. Riedel, P. Böhi, Y. Li, T. W. Hänsch, A. Sinatra,
and P. Treutlein, Nature 464, 1170 (2010).
[5] C. Gross, T. Zibold, E. Nicklas, J. Estève, and M. K.
Oberthaler, Nature 464, 1165 (2010).
[6] H.F.HofmannandS.Takeuchi, Phys.Rev.A 68, 032103
(2003).
[7] G. Tóth, C. Knapp, O. Gühne, and H. J. Briegel, Phys.
Rev. A 79, 042334 (2009).
[8] E. G. Cavalcanti, P. D. Drummond, H. A. Bachor, and
M. D. Reid, Opt. Express 17, 18693 (2009).
[9] M. D. Reid, P. D. Drummond, W. P. Bowen, E. G. Cav-
alcanti, P. K. Lam, H. A. Bachor, U. L. Andersen, and
G. Leuchs, Rev. Mod. Phys. 81, 1727 (2009).
[10] E. G. Cavalcanti, S. J. Jones, H. M. Wiseman, and M.
D. Reid, Phys. Rev. A 80, 032112 (2009).
[11] C. D. Hamley, C. S. Gerving, T. M. Hoang, E. M. Book-
jans, and M. S. Chapman, Nat. Phys. 8, 305 (2012).
[12] C. Chin, R. Grimm, P. Julienne, and E. Tiesinga, Rev.
Mod. Phys. 82, 1225 (2010).[13] Y. J. Lin, K. Jiménez-García, and I. B. Spielman, Nature
471, 83 (2011).
[14] P. Wang, Z. Q. Yu, Z. Fu, J. Miao, L. Huang, S. Chai, H.
Zhai, and J. Zhang, Phys. Rev. Lett. 109, 095301 (2012).
[15] L. W. Cheuk, A. T. Sommer, Z. Hadzibabic, T. Yefsah,
W. S. Bakr, and M. W. Zwierlein, Phys. Rev. Lett. 109,
095302 (2012).
[16] H. Zhai, Rep. Prog. Phys. 78, 026001 (2015).
[17] W. Yi, W. Zhang, and X. Cui, Sci. China: Phys. Mech.
Astron. 58, 014201 (2015).
[18] J. Zhang, H. Hu, X. J. Liu, and H. Pu, Annu. Rev. Cold
At. Mol. 2, 81 (2014).
[19] J. Lian, L. Yu, J.-Q. Liang, G. Chen, and S. Jia, Sci.
Rep.3, 3166 (2013).
[20] Y. Huang and Z.-D. Hu, Sci. Rep. 5, 8006 (2015).
[21] E. Yukawa, M. Ueda, and K. Nemoto, Phys. Rev. A 88,
033629 (2013).
[22] L. M. Duan, J. I. Cirac, and P. Zoller, Phys. Rev. A 65,
033619 (2002).
[23] Özgür E. Müstecaphoˇ glu, M. Zhang, and L. You, Phys.
Rev. A 66, 033611 (2002).
[24] Z. Lan and P. Ohberg, Phys. Rev. A 89, 023630 (2014).
[25] J. Ma, X. Wang, C. P. Sun, and F. Nori, Phys. Rep. 509,
89 (2011).
[26] W. Zhang, D. L. Zhou, M. S. Chang, M. S. Chapman,
and L. You, Phys. Rev. A 72, 013602 (2005). |
2201.11823v2.Superfluid_transition_temperature_and_fluctuation_theory_of_spin_orbit_and_Rabi_coupled_fermions_with_tunable_interactions.pdf | Super
uid transition temperature and
uctuation theory of
spin-orbit and Rabi coupled fermions with tunable interactions
Philip D. Powell,1, 2Gordon Baym,2and C. A. R. S a de Melo3
1Lawrence Livermore National Laboratory, 7000 East Avenue, Livermore, California 94550, USA
2Department of Physics, University of Illinois at Urbana-Champaign,
1110 W. Green Street, Urbana, Illinois 61801, USA
3School of Physics, Georgia Institute of Technology, 837 State Street, Atlanta, Georgia 30332, USA
(Dated: June 6, 2022)
We obtain the super
uid transition temperature of equal Rashba-Dresselhaus spin-orbit- and
Rabi-coupled Fermi super
uids, from the Bardeen-Cooper-Schrieer (BCS) to Bose-Einstein con-
densate (BEC) regimes in three dimensions for tunable s-wave interactions. In the presence of Rabi
coupling, we nd that spin-orbit coupling enhances (reduces) the critical temperature in the BEC
(BCS) limit. For xed interactions, we show that spin-orbit coupling can convert a rst-order (dis-
continuous) phase transition into a second-order (continuous) phase transition, as a function of Rabi
coupling. We derive the Ginzburg-Landau free energy to sixth power in the super
uid order pa-
rameter to describe both continuous and discontinuous phase transitions as a function of spin-orbit
and Rabi couplings. Lastly, we develop a time-dependent Ginzburg-Landau
uctuation theory for
an arbitrary mixture of Rashba and Dresselhaus spin-orbit couplings at any interaction strength.
I. INTRODUCTION
The ability to simulate magnetic and other external
elds [1{12] in cold atomic gases has created the oppor-
tunity to explore a wide variety of new interactions and
complex phase structures otherwise inaccessible in the
laboratory. Moreover, the capacity to generate these syn-
thetic elds in both bosonic and fermionic systems, and
to continuously tune two-body interactions by means of a
Feshbach resonance, has opened up a wonderland of tun-
able systems, previously restricted to theorists' dreams.
For example, the possibility of simulating quantum chro-
modynamics (QCD) on an optical lattice [13{16] is a tan-
talizing prospect for researchers whose current theoreti-
cal tools remain limited by QCD's non-perturbative char-
acter and the restriction of lattice techniques to near-zero
chemical potential.
Previous theoretical analyses of three-dimensional
spin-orbit-coupled Fermi gases (e.g.,6Li,40K) have fo-
cused mainly on the zero-temperature limit, in which sev-
eral exotic phases characterized by unconventional pair-
ing are expected to emerge [17{22]. However, the Ra-
man laser platforms currently employed to produce syn-
thetic spin-orbit elds also induce heating that prevents
the realization of temperatures suciently low to observe
the super
uid transition in either the weakly coupled
Bardeen-Cooper-Schrieer (BCS) or the strongly coupled
Bose-Einstein condensate (BEC) regimes [5, 7]. Thus,
while two-body bound states (Feshbach molecules) have
been observed in the BEC limit of40K [7, 23], the obser-
vation of super
uid states remain elusive. Future exper-
iments, however, may break this impasse by employing
a new platform currently under development|the radio-
frequency atom chip|which avoids heating of the atom
cloud entirely [24]. While rf atom chips are somewhat
more restricted than the Raman scheme in the maximum
obtainable spin-orbit coupling, its potential to reach su-per
uid temperatures is leading to its adoption in the
next generation of experiments probing the topological
super
uid phases of spin-orbit-coupled fermions [25].
One class of systems of particular interest in the con-
text of quantum simulation is that of Rashba-Dresselhaus
spin-orbit-coupled gases [17{22, 26, 27]. These systems
are intriguing both because they re
ect physics studied
extensively in the context of semiconductors [28, 29], and
because they provide a platform for realizing tunable
non-Abelian elds in the laboratory. Thus, while the
holy grail of a full optical simulation of QCD remains
years in the future, there do exist notable analogies be-
tween quark matter and cold atomic systems (e.g., non-
Abelian elds, evolution between strongly and weakly
coupled limits) within near-term experimental reach [30{
33]. Investigations of spin-orbit-coupled ultracold gases
have also included optical lattices [34{43], thus enlarg-
ing the number of possible physical systems that can be
accessible experimentally.
To date, most experimental realizations of these
systems have adopted equal Rashba-Dresselhaus cou-
plings [4{6, 44], but systems exhibiting Rashba-only cou-
plings have also been created [11, 45, 47]. Other ex-
periments have generated spin-orbit coupling dynami-
cally [48] or even created three-dimensional spin-orbit
coupling [49]. Due to the versatility of Rashba-
Dresselhaus coupled systems, the ability to realize these
systems in the laboratory, and the myriad technical
challenges inherent in reaching arbitrarily low temper-
atures, it is increasingly important to provide a theoret-
ical framework for guiding and testing these simulators
against experimental probes at realistic (nonzero) tem-
peratures.
This problem bears a close relation to spin-orbit cou-
pling in solids, where the role of the Rabi frequency is
played by an external Zeeman magnetic eld. While a
mean-eld treatment describes well the evolution from
the BCS to the BEC regime at zero temperature [50, 51],arXiv:2201.11823v2 [cond-mat.quant-gas] 2 Jun 20222
this order of approximation fails to describe the correct
critical temperature of the system in the BEC regime
because the physics of two-body bound states, i.e., Fes-
hbach molecules, is not captured when the pairing order
parameter goes to zero [52]. To remedy this problem, we
include the eects of order-parameter
uctuations in the
thermodynamic potential.
In this paper, we investigate the impact of a specic
class of spin-orbit coupling, namely, an equal mixture
of Rasha and Dresselhaus terms, on the super
uid tran-
sition temperature of a three-dimensional Rabi-coupled
Fermi gas, but also give general results for an arbitrary
mixture of Rashba and Dresselhaus components. This
paper is the longer version of our preliminary work [53].
We stress that the present results are applicable to both
neutral cold atomic and charged condensed-matter sys-
tems. We show that spin-orbit coupling, in the presence
of a Rabi eld (or Zeeman eld, in solids), enhances the
critical temperature of the super
uid in the BEC regime
and converts a discontinuous rst-order phase transition
into a continuous second-order transition, as a function of
the Rabi frequency for given two-body interactions. We
analyze the nature of the phase transition in terms of the
Ginzburg-Landau free energy, calculating it to the sixth
power of the super
uid order parameter, as required to
describe both discontinuous transitions as a function of
the spin-orbit coupling, Rabi frequency, and two-body
interactions.
This paper is organized as follows. In Sec. II, we de-
scribe the Hamiltonian and action for three-dimensional
Fermi gases in the presence of a general Rashba-
Dresselhaus spin-orbit coupling, Rabi eld, and tunable
s-wave interactions. We also obtain the inverse Green
operator that is used in the calculation of the thermo-
dynamic potential and Ginzburg-Landau theory of sub-
sequent sections. In Sec. III, we analyze the thermo-
dynamic potential across the entire BCS-to-BEC evolu-
tion, including contributions from both the mean-eld
and Gaussian
uctuations, and obtain the order parame-
ter and number equations. In Sec. IV, we study the com-
bined eects of Rabi elds and spin-orbit coupling on the
super
uid critical temperature, constructing the nite-
temperature phase diagram versus Rabi elds and scat-
tering parameter. In Sec. V, we present the Ginzburg-
Landau (GL) theory for the super
uid order parameter
and investigate further corrections to the critical tem-
perature in the BEC limit by including interactions be-
tween bosonic bound states. The GL action is obtained
to sixth order in the order parameter to allow for the exis-
tence of discontinuous (rst-order) phase transitions. In
Sec. VI, we compare our work on the experimentally rel-
evant equal Rashba-Dresselhaus spin-orbit coupling with
earlier work that has considered dierent forms of the-
oretically motivated spin-orbit couplings. In Sec. VII,
we conclude and look toward the future of experimental
work in this eld.
In the interest of readability, we relegate a number of
detailed calculations to appendices. In Appendix A, wediscuss the Hamiltonian and eective Lagrangian for a
general Rashba-Dresselhaus spin-orbit coupling. In Ap-
pendix B, we analyze the saddle-point approximation for
general Rashba-Dresselhaus spin-orbit coupling. In Ap-
pendix C, we derive the modied number equation, in-
cluding the contribution arising from Gaussian
uctua-
tions, which renormalizes the chemical potential obtained
at the saddle-point level. In Appendix D, using a gen-
eral Rashba-Dresselhaus spin-orbit coupling, we obtain
expressions for the coecients of the Ginzburg-Landau
theory up to sixth order in order parameter.
II. HAMILTONIAN AND ACTION
Throughout this paper, we adopt units in which ~=
kB= 1. The Hamiltonian density of a three-dimensional
Fermi gas in the presence of Rashba-Dresselhaus spin-
orbit coupling and Rabi eld is
H(r) =Hk(r) +Hso(r) +HI(r) n(r): (1)
The rst term in Eq. (1) is the kinetic energy,
Hk(r) =X
s y
s(r)^k2
2m s(r); (2)
where ^k= iris the momentum operator, s(r) is the
fermion eld at position rwith (real or pseudo-) spin s
and massm. The second term is the spin-orbit interac-
tion,
Hso(r) =X
ss0 y
s(r)h
Hso(^k)i
ss0 s0(r); (3)
with the spin-orbit coupling matrix in momentum ( k)
space being
Hso(^k) =
m(^kxx+^kyy)
R
2z; (4)
where (x;y;z) are the Pauli matrices in spin space,
is the momentum transfer to the atoms in a two-photon
Raman process [7] or on a radio frequency atom chip [24],
is the anisotropy of the Rashba-Dresselhaus eld, and
Ris the Rabi frequency. The third term is the two-body
s-wave contact interaction,
HI(r) = g y
"(r) y
#(r) #(r) "(r); (5)
whereg>0 corresponds to a constant attraction between
opposite spins. Finally, is the chemical potential and
n(r) =P
s y
s(r) s(r) is the local density. While the gen-
eral Rashba-Dresselhaus spin-orbit coupling is discussed
in Appendix A, in what follows we focus on the more
experimentally relevant situation of equal Rashba and
Dresselhaus couplings ( = 0).
Standard manipulations (see Appendix A) lead to the
Lagrangian density,
L(r;) =1
2 y(r;)G 1(^k;) (r;) +1
gj(r;)j2
+K(^k)(r r0); (6)3
where=itis the imaginary time, = ( " # y
" y
#)T
is the Nambu spinor, K(^k) =^k2=2m is the kinetic
energy operator with respect to the chemical potential,
and ( r;) = gh #(r;) "(r;)iis the pairing eld
describing the formation of pairs of two fermions with
opposite spins. Note that includes the overall positive
shift2=2min the single-particle kinetic energies due to
spin-orbit coupling. The inverse Green's operator ap-
pearing in Eq. (6) is
G 1(^k;) =0
BB@@ K" i^kx=m 0
i^kx=m @ K# 0
0 @+K" i^kx=m
0i^kx=m @+K#1
CCA;
(7)
whereK";#=K(^k)
R=2;are the kinetic energy terms
shifted by the Rabi coupling.
As noted above, a mean-eld treatment of this La-
grangian fails to correctly describe the super
uid critical
temperature in the BEC regime. However, the inclu-
sion of Gaussian
uctuations of captures the eects of
two-body bound states and leads to a physical super
uid
transition temperature. It is to this task that we now
turn.
III. THERMODYNAMIC POTENTIAL
The system's partition function may be expressed in
terms of the functional integral,
Z=Z
DDD D ye S; (8)
where the Euclidean action is
S=Z
0dZ
d3rL(r;); (9)
= 1=Tis the inverse temperature, and the Lagrangian
density is given by Eq. (6). Integrating over the fermion
elds yields the thermodynamic potential,
= TlnZ=
0+
F; (10)
where
0= TlnZ0=TS0is the mean-eld (saddle-
point) contribution, for which ( r;) = 0, and
F= TlnZFis the contribution arising from order-
parameter
uctuations. Detailed derivations of the ther-
modynamic potential for a general Rashba-Dresselhaus
spin-orbit coupling, as well as the associated order pa-
rameter and number equations, are given in Appen-
dices B and C. The contributions to the thermody-
namic potential for the experimentally relevant situa-
tion of equal Rashba-Dresselhaus spin-orbit coupling are
discussed below in Sec. III A at the mean-eld and in
Sec. III B at the Gaussian
uctuation level.A. Mean-Field Approximation
The mean-eld, or saddle-point, term in the thermo-
dynamic potential is
0=Vj0j2
g T
2X
k;jlnh
1 +e Ej(k)i
+X
kk;(11)
wherek="k ,"k=k2=2m, and theEj(k), with
j=f1;2;3;4g, are the eigenvalues of the momentum
space Nambu Hamiltonian matrix,
H0(k) =@ G 1(k;)j= 0; (12)
where the operator @=I@, andIis the identity matrix.
The rst set of eigenvalues,
E1;2(k) =2
42
k2s
E2
0;kh2
k kx
m2
j0j23
51=2
(13)
describe quasiparticle excitations, with the plus (+) as-
sociated with E1and the minus ( ) withE2. The
second set of eigenvalues, E3;4(k) = E2;1(k);corre-
sponds to quasiholes. Further, 2
k=E2
0;k+h2
k, where
E0;k=p
2
k+j0j2;andhk=p
(kx=m)2+
2
R=4 is
the magnitude of the combined spin-orbit and Rabi cou-
plings.
We express the two-body interaction parameter gin
terms of the renormalized s-wave scattering length as
via the relation [52]
1
g= m
4as+1
VX
k1
2"k: (14)
Note thatasis thes-wave scattering length in the absence
of spin-orbit and Rabi elds. It is, of course, possible to
expressg, and all subsequent relations, in terms of a scat-
tering length which is renormalized by the presence of the
spin-orbit and Rabi elds [54, 55], but for both simplic-
ity and the sake of referring to the more experimentally
accessible quantity, we do not do so here.
The order-parameter equation is obtained from the
saddle-point condition
0=
0jT;V; = 0, leading to
m
4as=1
2VX
k1
"k A+(k)
2
R
4khkA (k)
;(15)
where we introduced the notation
A(k) =1 2n1(k)
2E1(k)1 2n2(k)
2E2(k); (16)
withnj(k) = 1=
eEj(k)+ 1
being the Fermi function.
In addition, the particle number at the saddle point N0=
@
0=@jT;V;is given by
N0=X
k
1 k
A+(k) +(kx=m)2
khkA (k)
:(17)4
The mean-eld temperature T0is determined by solv-
ing Eq. (15) for the given . The corresponding number
of particles is given by Eq. (17). This mean-eld treat-
ment leads to a transition temperature e1=kFas, where
kFis the Fermi momentum. This result gives the correct
transition temperature on the BCS limit; however, it is
unphysical on the BEC regime for kFas!0. In order to
nd a physical result, we need to include order-parameter
uctuations, which we now do.
B. Gaussian Fluctuations
In discussing Gaussian
uctuations, we concentrate on
equal Rasha-Dresselhaus couplings, leaving details for
general Rashba-Dresselhaus coupling to Appendix C.
To obtain the correct super
uid transition temperature
in the BEC limit we must include the physics of two-body
bound states near the transition, as described by the two-
particleT-matrix [56, 57]. Accounting for all two-particle
channels, the T-matrix calculation leads to a two-particle
scattering amplitude , where
1(q;z) =m
4as 1
2VX
k1
"k+2X
i;j=1ijWij
; (18)
zis the complex frequency and
Wij=1 ni(k) nj(k+q)
z Ei(k) Ej(k+q): (19)
At the super
uid phase boundary 0!0, the eigenval-
ues appearing in Eq. (19) reduce to E1;2(k) =jjkjhkj,
but it is straightforward to show that ignoring the abso-
lute values does not result in any change in either the
mean-eld order parameter or number equation. Mean-
while, the coecients
11=22=jukuk+q vkv
k+qj2; (20)
12=21=jukvk+q+uk+qvkj2; (21)
are the coherence factors associated with the quasi-
particle amplitudes for 0= 0:
uk=s
1
2
1 +
R
2hk
; v k=is
1
2
1
R
2hk
:(22)
The Gaussian
uctuation correction to the thermody-
namic potential is
F= TX
q;iqnln [ (q;iqn)=V]: (23)
over the entire BCS-to-BEC evolution. The
uctuation
contribution to the particle number is therefore NF=
@
F=@jT;V, where
NF=X
qZ1
1d!
nB(!)@(q;!)
@ @(q;0)
@
T;V;
(24)with the phase shift (q;!) dened via the relation
(q;!i) =j (q;!)jei(q;!): (25)
When two-body states are present, the
uctuation con-
tribution can be written as NF=Nsc+Nb, where
Nsc=X
qZ1
!tp(q)d!
nB(!)@(q;!)
@ @(q;0)
@
T;V
(26)
is the number of particles in scattering states, and !tp(q)
is the two-particle continuum threshold corresponding to
the branch point of 1(q;z) [56, 58],
Nb= 2X
qnB(Ebs(q) 2); (27)
is the number of fermions in bound states, where
nB(!) = 1=(e! 1) is the Bose distribution function,
andEbs(q) is the energy of the bound states obtained
from 1(q;z=E 2) = 0, corresponding to a pole
in the scattering amplitude ( q;z). In the limit of large
and negative fermion chemical potential, the system be-
comes non-degenerate and 1(q;z) = 0 becomes the
exact eigenvalue equation for the two-body bound state
in the presence of spin-orbit and Rabi coupling [23]. The
total number of fermions, as a function of , thus be-
comes
N=N0+NF; (28)
whereN0is given in Eq. (17) and NFis the sum of Nsc
andNb, as discussed above [52, 56].
IV. CRITICAL TEMPERATURE
We calculate numerically the transition temperature
Tcbetween the normal and uniform super
uid states, as
a function of the scattering parameter 1 =kFas, by simul-
taneously solving the order parameter and number equa-
tions (15) and (28). The solutions correspond to the min-
ima of the free energy, F=
+N. We do not discuss
the cases of Fulde-Ferrell [59] or Larkin-Ovchinnikov [60]
nonuniform super
uid phases since they only exist over a
very narrow region of the phase diagram deep in the BCS
regime [59, 60], which is not experimentally accessible for
ultracold fermions.
Figure 1, in which we scale temperatures by the Fermi
temperature TF=k2
F=2m, shows the eects of spin-
orbit and Rabi couplings on Tc. The solid (black) line
in Fig. 1(a) shows Tcversus 1=kFasfor zero Rabi cou-
pling (
R= 0) and zero spin-orbit coupling . If
R= 0, the spin-orbit coupling can be removed by a
simple gauge transformation, and thus plays no role. In
this situation, the pairing is purely s-wave. The dashed
(blue) line shows Tcfor
R6= 0, with vanishing equal
Rashba-Dresselhaus spin-orbit coupling. We see that for
xed interaction strength, the pair-breaking eect of the5
FIG. 1. (Color online) (a) The super
uid transition tem-
peratureTc=TF, whereTFis the Fermi temperature, vs the
scattering parameter 1 =kFasfor equal Rashba-Dresselhaus
spin-orbit coupling and two dierent Rabi coupling strengths,
R= 0 and"F. For
R= 0 [solid (black) curve], Tcis the
same as for zero spin-orbit coupling since the equal spin-orbit
eld can be gauged away. The dashed (blue) line shows Tc
for zero spin-orbit coupling, with
R="F, while the dot-
ted (green) line shows Tcfor
R="Fand ~==kF= 0:5.
(b)Tcis drawn at unitarity, 1 =kFas= 0, and in the inset
at 1=kFas= 2:0, as a function of e
R=
R="F. The solid
(red) curves represent ~ = 0 and the dotted (blue) curves
represent ~= 0:5. Across the dotted (red) curves, the phase
transition is rst order.
Rabi coupling suppresses super
uidity, compared with
R= 0; the Rabi eld here plays the pair-breaking role
of the Zeeman eld in an superconductor.
With both spin-orbit and Rabi couplings present, the
two-particle pairing is no longer purely singlet s-wave,
but obtains a triplet p-wave component; the admixture
stabilizes the super
uid phase, as shown by the dotted
(green) line. The latter curve shows that in the BEC
regime with large positive 1 =kFas, the super
uid tran-
sition temperature is enhanced by the presence of spin-
orbit and Rabi couplings, a consequence of the reduction
FIG. 2. (Color online) Chemical potential at the super
uid
critical temperature ( Tc) for ~==kF= 0:5 and various Rabi
elds, e
R=
R="F.
of the bosonic eective mass in the xdirection below 2 m.
However, for suciently large
R, the geometric mean
bosonic mass MBincreases above 2 mandTcdecreases.
This renormalization of the mass of the bosons can be
traced back to a change in the energy dispersion of the
fermions when both spin-orbit coupling and Rabi elds
are present.
Figure 1(b) shows Tcversus
Rfor xed 1=kFas, both
with and without equal Rashba-Dresselhaus spin-orbit
coupling at = 0:5kF. When both andTare zero,
super
uidity is destroyed at a critical value of
Rcorre-
sponding to the Clogston limit [61]. At low temperature,
the phase transition to the normal state is rst order
because the Rabi coupling is suciently large to break
singlet Cooper pairs. However, at higher temperatures
the singlet s-wave super
uid starts to become polarized
by thermally excited quasiparticles that produce a para-
magnetic response. Thus, above the characteristic tem-
perature indicated by the large (red) dots, the transition
becomes second order, as pointed out by Sarma [62]. The
change in the transition order occurs not only for = 0,
but also for nonzero values of both in the BCS regime
and near unitarity, depending on the choice of parame-
ters, as illustrated in Fig. 1(b).
The critical temperature for 6= 0 vanishes only
asymptotically in the limit of large
R. We note that
for
R=EFand= 0, the transition from the super-
uid to the normal state is continuous at unitarity, but
very close to a discontinuous transition. In the range
1:05.
R=EF.1:10, numerical uncertainties as !0
prevent us from predicting exactly whether the transition
at unitarity is continuous or discontinuous.
Figure 2 shows (Tc) for xed spin-orbit coupling and
several Rabi couplings. The solid (black) curve, which
represents the situation in which no Rabi eld is present,
is equivalent to the situation in which spin-orbit coupling
is also absent, as noted in the discussion of Fig. 1. It is6
FIG. 3. (Color online) Phase diagram of critical temperature
Tc=TFvs 1=kFasand
R="Ffor equal Rashba-Dresselhaus
coupling=kF= 0:5. The nite-temperature uniform super-
uid phases re
ect those at T= 0 shown in the background.
These phases are distinguished by the number of rings (line
nodes) in the quasiparticle excitation spectrum [i.e., where
E2(k) = 0] and type of gap: (1) direct gapped super
uid with
zero rings (magenta diamonds), (2) indirect gapped super
uid
with zero rings (red circles), (3) gapless super
uid with two
rings (blue square), and (4) gapless one-ring super
uid (green
stars).
evident that while the Rabi eld reduces the chemical
potential in the BCS limit, it also shifts the onset of the
system's evolution to the BEC limit to larger inverse scat-
tering lengths, and produces a non-monotonic behavior
of(Tc) near unitarity.
Figure 3 shows Tcfor equal Rashba-Dresselhaus cou-
pling= 0:5kF, as a function of Rabi eld and
scattering parameter. We also superpose the zero-
temperature phase diagram to illustrate the dierent su-
per
uid ground states of this system. According to the
zeros of the lowest quasiparticle energy E2(k), the uni-
form super
uid phases that emerge are [21] direct gapped
with zero rings (line nodes), indirectly gapped with zero
rings, gapless with one ring, and gapless with two rings.
Figure 4 shows the fractional number Nb=Nof bound
fermions at Tcas a function of 1 =kFasfor two sets of
external elds. In the BCS (BEC) regime, the rela-
tive contribution to Nis dominated by unbound (bound)
fermions. The main eect of spin-orbit and Rabi elds on
Nb=Nis to shift the location where the two-body bound
states emerge. For xed spin-orbit coupling (Rabi eld)
and increasing Rabi eld (spin-orbit coupling), two-body
bound states emerge at larger (smaller) scattering pa-
rameters. These shifts are in agreement with the cal-
culated shifts in binding energies of Feshbach molecules
in the presence of equal Rashba-Dresselhaus spin-orbit
coupling and Rabi elds [23].
FIG. 4. (Color online) Fractional number Nb=Nof bound
fermions as a function of the interaction parameter 1 =kFas,
for equal Rashba-Dresselhaus coupling =kF= 0:5 and Rabi
frequencies e
R=
R="F= 0 (black solid line) and e
R=
R="F= 2 (red dot-dashed line).
V. GINZBURG-LANDAU THEORY
To further elucidate the eects of
uctuations on the
order of the super
uid transition, as well as to assess the
impact of spin-orbit and Rabi couplings near the crit-
ical temperature, we now derive the Ginzburg-Landau
description of the free energy near the transition. In the
limit of small order parameter, the
uctuation action SF
can be expanded in powers of the order parameter ( q)
beyond Gaussian order. The expansion of SFto quar-
tic order is sucient to describe the continuous (second-
order) transition in Tcversus 1=kFasin the absence of a
Rabi eld [52]. However, to correctly describe the rst-
order transition [61, 62] at low temperature (Fig. 1), it is
necessary to expand the free energy to sixth order in .
The quadratic (Gaussian-order) term in the action is
SG=VX
qjqj2
(q;z): (29)
For an order parameter varying slowly in space and time,
we may expand 1as
1(q;z) =a+X
`c`q2
`
2m d0z+; (30)
with the sum over `=fx;y;zg. The full result, as a
functional of ( r;), has the form
SF=Z
0dZ
d3r
d0@
@ +ajj2
+X
`c`jr`j2
2m+b
2jj4+f
3jj6
:(31)
The full time-dependent Ginzburg-Landau action de-
scribes systems in and near equilibrium (e.g., with col-7
lective modes). The imaginary part of d0measures the
non-conservation of jj2in time (i.e., the Cooper pair
lifetime). Details of the derivation of SFare found in
Appendix D.
We are interested in systems at thermodynamic equi-
librium, where the order parameter is independent of
time, that is, ( r;) = ( r). In this situation, mini-
mizing the free energy TSFwith respect to yields the
Ginzburg-Landau equation,
X
`c`r2
`
2m+bj(r)j2+fj(r)j4+a!
(r) = 0:
(32)
Forb > 0, the system undergoes a continuous phase
transition when achanges sign. However, when b <0,
the system is unstable in the absence of f. Forb <0
anda > 0, a rst-order phase transition occurs when
3b2= 16af. Positivefstabilizes the system even when
b<0.
In the BEC regime, where d0is purely real, we dene
an eective bosonic wave function ( r) =pd0(r) to re-
cast Eq. (32) in the form of the Gross-Pitaevskii equation
for a dilute Bose gas,
X
`r2
`
2M`+U2j (r)j2+U3j (r)j4 B!
(r) = 0:
(33)
Here,B= a=d0is the bosonic chemical potential,
M`=m(d0=c`) are the anisotropic bosonic masses, and
U2=b=d2
0andU3=f=d3
0represent contact interac-
tions of two and three bosons. In the BEC regime,
these terms are always positive, leading to a dilute gas
of stable bosons. The boson chemical potential Bis
2+Eb<0, whereEb= Ebs(q=0) is the two-
body binding energy in the presence of spin-orbit cou-
pling and Rabi frequency, obtained from the condition
1(q;E 2) = 0, discussed earlier.
The anisotropy of the eective bosonic masses, Mx6=
My=MzM?, stems from the anisotropy of the equal
Rashba-Dresselhaus spin-orbit coupling, which together
with the Rabi coupling modies the dispersion of the
constituent fermions along the xdirection. In the limit
kFas1, the many-body eective masses reduce to
those obtained by expanding the two-body binding en-
ergy,Ebs(q) Eb+P
`q2
`=2M`;and agree with known
results [23]. However, for 1 =kFas.2, many-body and
thermal eects produce deviations from the two-body re-
sult.
In the absence of two- and three-body boson-boson
interactions, U2andU3, we directly obtain an analytic
expression for Tcin the Bose limit from Eq. (27),
Tc=2
MBnB
(3=2)2=3
; (34)
withMB= (MxM2
?)1=3, by noting that B= 0 or
Ebs(q=0) 2= 0, and using the condition that nB'n=2 [with corrections exponentially small in (1 =kFas)2],
wherenBis the density of bosons. In the BEC regime,
the results shown in Fig. 1 include the eects of the mass
anisotropy, but do not include the eects of boson-boson
interactions.
To account for boson-boson interactions, we adopt the
Hamiltonian of Eq. (33) with U26= 0, but with U3= 0,
and apply the method developed in Ref. [63] to show that
these interactions further increase TBEC to
Tc(aB) = (1 +
)TBEC; (35)
where
=n1=3
BaB. Here,aBis thes-wave boson-
boson scattering length, is a dimensionless constant
1, and we use the relation U2= 4aB=MB. Since
nB=k3
F=62and the boson-boson scattering length is
aB=U2MB=4, we have
=~fMBeU2;wherefMB=
MB=2m;eU2=U2k3
F="F;and~==4(65)1=3=50:
For xed 1=kFas,Tcis enhanced by the spin-orbit eld,
a
R-dependent decrease in the eective boson mass MB
(10-15%), as well as a stabilizing boson-boson repulsion
U2(2-3%), for the parameters used in Fig. 1.
In closing our discussion of the strongly bound BEC
limit, we note that in the absence of spin-orbit coupling,
a Gaussian-order calculation of the two-boson scattering
length yields the erroneous Born approximation result
aB= 2as. However, an analysis of the T-matrix beyond
Gaussian order, which includes the eects of two-body
bound states, obtains the correct result aB= 0:6asat
very low densities [64] and agrees with four-body calcu-
lations [65]. The same method can be used to estimate U2
oraBbeyond the Born approximation discussed above.
Nevertheless, while the precise quantitative relation be-
tweenaBandasin the presence of spin-orbit coupling is
yet unknown, the trend of increasing Tcdue to spin-orbit
coupling has been clearly shown.
VI. COMPARISON TO EARLIER WORK
In this section, we brie
y compare our results with ear-
lier investigations of dierent types of theoretically mo-
tivated spin-orbit couplings, worked in dierent dimen-
sions, or at zero temperature. Our results focus mainly
on an analysis of the critical super
uid temperature and
the eects thereon of order-parameter
uctuations for
a three-dimensional Fermi gas in the presence of equal
Rashba-Dresselhaus spin-orbit coupling and Rabi elds.
The appendices consider the more general situation of
arbitrary Rashba and Dresselhaus components.
Several works have analyzed the eects of spin-orbit-
coupled fermions in three dimensions at zero tempera-
ture [17{22, 66{69]. While some authors have described
the situation of Rashba-only couplings [17{19, 66], others
have assessed the case of equal Rashba and Dresselhaus
components [21, 22] or a general mixture of the two [20].
It has been demonstrated that in the absence of a Rabi
eld, the zero-temperature evolution from BCS to BEC8
super
uidity is a crossover for s-wave systems, not only
for Rashba-only couplings [17{20, 66], but also for ar-
bitrary Rashba and Dresselhaus components [20]. This
result directly follows from the fact that the quasiparti-
cle excitation spectrum remains fully gapped throughout
the evolution.
In contrast, the addition of a Rabi eld gives rise
to topological phase transitions for Rashba-only cou-
plings [17] and equal Rashba and Dresselhaus compo-
nents [21, 22], a situation which certainly persists for
general Rashba-Dresselhaus couplings. The simultane-
ous presence of a general Rashba-Dresselhaus spin-orbit
coupling and Rabi elds leads to a qualitative change in
the quasiparticle excitation spectrum and to the emer-
gence of topological super
uid phases [17, 21, 22]. Two-
dimensional systems have also been investigated at zero
temperature, where topological phase transitions have
been identied for Rashba-only [70] and equal Rashba-
Dresselhaus [71] couplings, in the presence of a Rabi eld.
While early papers in this eld focused mainly
on the zero-temperature limit, progress toward nite-
temperature theories was made rst in two dimen-
sions [72, 73] and later in three dimensions [74{76].
The eects of a general Rashba-Dresselhaus spin-orbit
coupling and Rabi eld on the Berezenskii-Kosterlitz-
Thouless transition were thoroughly investigated for
two-dimensional Fermi gases at nite temperatures [72,
73], including both Rashba-only and equal Rashba-
Dresselhaus spin-orbit couplings as examples.
The super
uid critical temperature in three dimen-
sions was investigated using a spherical (3D) spin-orbit
couplingkin the absence of a Rabi eld [74, 75],
and also for Rashba-only (2D) couplings in the pres-
ence of a Rabi eld [76]. In a recent review article [77],
the critical temperature throughout the BCS-BEC evo-
lution was discussed both in the absence [52] and pres-
ence [53] of Rashba-Dresselhaus spin-orbit coupling. In
Secs. 5 and 6 of this review, the authors describe the
same method and expressions we obtained in our earlier
preliminary work [53] for the analytical relations required
to obtain the critical temperature at the Gaussian order;
they include, however, only the contribution of bound
states discussed earlier in the literature for Rashba-only
spin-orbit coupling without Rabi elds [18]. In contrast,
here we develop a complete Gaussian theory to compute
the super
uid critical temperature of a three-dimensional
Fermi gas in the presence of both a general Rashba-
Dresselhaus (2D) spin-orbit coupling and Rabi elds. We
focus our numerical calculations on the specic situation
of equal Rashba-Dresselhaus components, which is eas-
ier to achieve experimentally in the context of ultracold
atoms. Our key results, already announced in our earlier
work [53], include the contributions of bound and scat-
tering states at the Gaussian level. As seen in Fig. 4
of this present paper, there is a wide region of interac-
tion parameters for which the contribution of scattering
states cannot be neglected. Furthermore, unlike previ-
ous work [74{77], we provide a comprehensive analysisof the Ginzburg-Landau
uctuation theory and include
the eects of boson-boson interactions on the super
uid
critical temperature in the BEC regime.
VII. CONCLUSION
We have evaluated the super
uid critical tempera-
ture throughout the BCS-to-BEC evolution of three-
dimensional Fermi gases in the presence of equal Rashba-
Dresselhaus spin-orbit couplings, Rabi elds, and tun-
ables-wave interactions. Furthermore, we have devel-
oped the Ginzburg-Landau theory up to sixth power in
the order parameter to elucidate the origin of rst-order
phase transitions when the spin-orbit eld is absent and
the Rabi eld is suciently large. Lastly, in the appen-
dices, we have presented the nite-temperature theory of
s-wave interacting fermions in the presence of a generic
Rashba-Dresselhaus coupling and external Rabi elds,
as well as the corresponding time-dependent Ginzburd-
Landau theory near the super
uid critical temperature.
ACKNOWLEDGMENTS
We thank I. B. Spielman for discussions. The re-
search of P.D.P. was supported in part by NSF Grant
No. PHY1305891 and that of G.B. by NSF Grants
No. PHY1305891 and No. PHY1714042. Both G.B.
and C.A.R. SdM. thank the Aspen Center for Physics,
supported by NSF Grants No. PHY1066292 and No.
PHY1607611, where part of this work was done. This
work was performed under the auspices of the U.S. De-
partment of Energy by Lawrence Livermore National
Laboratory under Contract No. DE-AC52-07NA27344.
Appendix A: Hamiltonian and eective Lagrangian
for general Rashba-Dresselhaus spin-orbit coupling
In this appendix, we consider a larger class of spin-
coupled fermions in three dimensions with a general
Rashba-Dresselhaus (GRD) coupling. The Hamiltoninan
density for equal Rashba-Dresselhaus (ERD) discussed
in Sec. II is a particular case of the general Rashba-
Dresselhaus Hamiltonian density,
H(r) =H0(r) +Hso(r) +HI(r): (A1)
Adopting units in which ~=kB= 1, the independent-
particle Hamiltonian density without spin-orbit coupling
is
H0(r) =X
jr (r)j2
2m y
(r) (r)
;(A2)
where ,m, andare the fermion eld operator,
mass, and chemical potentials for internal state , re-
spectively. The spin-orbit Hamiltonian can be written9
as
Hso(r) = X
i y
(r)i;hi(r) (r); (A3)
where theiare the Pauli matrices in isospin (internal
state) space and h= (hx;hy;hz) includes both the spin-
orbit coupling and Zeeman elds. Finally, we consider a
two-bodys-wave contact interaction,
HI(r) = g y
"(r) y
#(r) #(r) "(r); (A4)
whereg>0 corresponds to an attractive interaction.
By introducing the pairing eld ( r;) =
gh #(r;) "(r;)i;we remove the quartic inter-
action and obtain the Lagrangian density,
L(r;) =1
2 y(r;)G 1(^k;) (r;) +j(r;)j2
g
+eK+(^k)(r r0); (A5)
where we introduced the momentum operator ^k= ir,
the Nambu spinor = ( " # y
" y
#)T, and dened
eK= (eK"eK#)=2:Here,eK"=K" hz;andeK#=
K#+hz;withK(^k) =^k2=(2m) being the kinetic
energy operator of internal state with respect to its
chemical potential. Lastly, the inverse Green's operator
appearing in Eq. (A5) is
G 1(^k;) =0
BB@@ eK"h
? 0
h?@ eK# 0
0 @+eK" h?
0 h
?@+eK#1
CCA;
(A6)
whereh?(^k) =hx(^k) +ihy(^k) plays the role of the spin-
orbit coupling, and hzis the Zeeman eld along the z
direction.
To make progress, we expand the order parameter
about its saddle-point (mean-eld) value 0by writ-
ing ( r;) = 0+(r;):Next, we integrate over the
fermionic elds and use the decomposition G 1(^k;) =
G 1
0(^k;)+G 1
F(^k;);where G 1
0(^k;) is the mean-eld
Green's operator, given by Eq. (A6) with ( r;) = 0,
andG 1
F(^k;) is the contribution to the inverse Green's
operator arising from
uctuations. These steps yield the
saddle-point Lagrangian density,
L0(r;) = T
2VTr ln(G 1
0) +j0j2
g+eK+(^k)(r r0);
(A7)
and the
uctuation contribution,
LF(r;) = T
2VTr ln( I+G0G 1
F) + ( r;) +j(r;)j2
g;
(A8)
resulting in the eective Lagrangian density Le(r;) =
L0(r;) +LF(r;):In the expressions above, we workin a volume Vand take traces over both discrete and
continuous indices. Notice that the term ( r;) =
[0(r;) +
0(r;)]=gin the
uctuation Lagrangian
cancels out the linear terms in andwhen the loga-
rithm is expanded, due to the saddle point condition
S0
0= 0; (A9)
whereS0=R
0dd3rL0(r;) is the saddle-point action.
Appendix B: Saddle Point Approximation for
general Rashba-Dresselhaus spin-orbit coupling
We rst analyze the saddle-point contribution. The
saddle-point thermodynamic potential
0= TlnZ0
can be obtained for the saddle-point partition function
Z=e S0as
0=TS0. Transforming the saddle-point
LagrangianL0from Eq. (A7) into momentum space and
integrating over spatial coordinates and imaginary time
leads to the saddle-point thermodynamic potential,
0=Vj0j2
g T
2X
k;jln(1+e Ek;j)+X
keK+(k);(B1)
whereK(k) =k2=2m and the eigenvalues Ek;j
are the poles of G0(k;z), withj=f1;2;3;4g.
Next, we restrict our analysis to mass balanced sys-
tems (m"=m#) in diusive equilibrium ( "=#).
We also consider the general Rashba-Dresselhaus (GRD)
spin-orbit eld h?(k) =(kx+iky)=m;whereand
are the magnitude and anisotropy of the spin-orbit cou-
pling, respectively. Note that this form is equivalent to
another common form of the Rashba-Dresselhaus cou-
pling found in the literature [21, 22]: hso=hR+hD
where hR=vR(kx^y ky^x) and hD=vD(kx^y+ky^x).
The two forms are related via a momentum-space ro-
tation and the correspondences =m(vR+vD) and
= (vR vD)=(vR+vD). The equal Rashba-Dresselhaus
limit (ERD) corresponds to vR=vD=v, leading to
= 0 and= 2mv. The specic case of equal Rashba-
Dresselhaus spin-orbit coupling discussed in the main
part of the paper corresponds to the case where = 0,
that is,h?(k) =kx=m:
For the general Rashba-Dresselhaus case, the four
eigenvalues are
E1;2(k) =h
2
k2q
E2
0;kh2
k j0j2jh?(k)j2i1=2
;(B2)
E3;4(k) = E2;1(k); (B3)
where the + ( ) sign within the outermost square root
corresponds to E1(E2), and the functions inside the
square roots are 2
k=E2
0;k+h2
k, with contributions
E0;k=q
2
k+j0j2; (B4)
hk=p
jh?(k)j2+h2z; (B5)10
wherek="k ;and"k=k2=2m:The order-
parameter equation is found from the saddle point con-
dition
0=
0jT;V; = 0. At the phase boundary be-
tween the super
uid and normal phases, 0!0, and
the order-parameter equation becomes
m
4as=1
2VX
k1
"k tanh(E1=2)
2E1 tanh(E2=2)
2E2
h2
z
khktanh(E1=2)
2E1 tanh(E2=2)
2E2
;
(B6)
after expressing the interaction parameter gin terms of
thes-wave scattering length via the relation
1
g= m
4as+1
VX
k1
2"k: (B7)
We note that asis thes-wave scattering length in the
absence of spin-orbit and Zeeman elds. It is, of course,
possible to express all relations obtained in terms of a
scattering length which is renormalized by the presence
of the spin-orbit and Rabi elds [54, 55]. However, in ad-
dition to complicating our already cumbersome expres-
sions, it would make reference to a quantity that is more
dicult to measure experimentally and that would hide
the explicit dependence of the properties that we analyze
in terms of the spin-orbit and Rabi elds, so we do not
consider such complications here. Note that since 0= 0
at the phase boundary, the eigenvalues in Eq. (B2) re-
duce toE1(k) =jkj+hk,E2(k) =jkj hk, which is
the absolute value of the normal-state energy dispersions.
However, it is straightforward to show that ignoring the
absolute values does not result in any change in either the
mean-eld order parameter given by Eq. (B6) or number
equation shown in Eq. (B8), when 0!0.
The saddle-point critical temperature T0is determined
by solving Eq. (B6) subject to the thermodynamic con-
straintN0= @
0=@jT;V;which yields
N0=X
k
1 k1
"k+tanh(E1=2)
2E1+tanh(E2=2)
2E2
+jh?(k)j2
khktanh(E1=2)
2E1 tanh(E2=2)
2E2
:
(B8)
A mean-eld description of the system, which involves
a simultaneous solution of Eqs. (B6) and (B8), yields
the asymptotically correct description of the system
in the BCS limit; however, such a description fails
miserably in the BEC regime where it does not ac-
count for the formation of two-body bound states.
The general Rashba-Dresselhaus spin-orbit saddle-point
equations (B6) and (B8) reduce to the equal Rashba-
Dresselhaus equations (15) and (17) of the main part
of the paper with the explicit use of hz=
R=2 and
h?(k) =kx=m, where
Ris the Rabi coupling.Appendix C: Derivation of the modied number
equation with Gaussian
uctuations
We begin by deriving the modied number equation
arising from Gaussian
uctuations of the order parame-
ter near the super
uid phase boundary. The
uctuation
thermodynamic potential
Fresults from the Gaussian
integration of the elds (r;) and(r;) in the
uctua-
tion partition function ZF=R
dde SF, where the ac-
tionSF=R
d
0R
d3rLF(r;) is calculated to quadratic
order. The contribution to the thermodynamic potential
due to Gaussian
uctuations is
F= TX
iqn;qln [ (q;iqn)=V] (C1)
whereqn= 2nT are the bosonic Matsubara frequencies
and ( q;iqn) is directly related to the pair
uctuation
propagator pair(q;iqn) =V 1(q;iqn):
The Matsubara sum can be evaluated via contour in-
tegration,
F= TX
qI
Cdz
2inB(z) ln [ (q;z)=V]; (C2)
wherenB(z) = 1=(ez 1) is the Bose function and the
countourCencloses all of the Matsubara poles of the
Bose function. Next, we deform the contour around the
Matsubara frequencies towards innity, taking into ac-
count the branch cut and the possibility of poles coming
from the logarithmic term inside the countour integral.
We take the branch cut to be along the real axis, then
add and subtract the pole at iqn= 0 to obtain
F= TX
qZ1
1d!
nB(!) [(q;!) (q;0)];(C3)
where the phase shift (q;!) is dened via ( q;!i) =
j (q;!)jei(q;!);and arises from the contour segments
above and below the real axis.
The thermodynamic identity N= @
=@jT;Vthen
yields to the
uctuation correction,
NF=TX
qZ1
1d!
nB(!)@(q;!)
@ @(q;0)
@
;
(C4)
to the the saddle-point number equation, and has a sim-
ilar analytical structure as in the case without spin-orbit
and Zeeman elds [52, 56]. Thus, we can write the -
nal number equation at the critical temperature Tcas
N=N0+NF. Since the phase shift (q;z) vanishes ev-
erywhere that ( q;z) is analytic, the only contributions
to Eq. (C4) arise from a possible isolated pole at !p(q)
and a branch cut extending from the two-particle contin-
uum threshold !tp(q) = minfi;j;kg[Ei(k) +Ej(k+q)] to
z!1 along the positive real axis. The explicit form of
(q;z) can be extracted from Eq. (D15) of Appendix D.
When there is a pole corresponding to the emergence of
a two-body bound state, we can explicitly write ( q;z)11
R(q)=(z !p(q));from which we obtain @(q;!)=@=
2(z !p(q));leading to the bound state density
Nb= 2X
qnB(!p(q)); (C5)
where the energy !p(q) must lie below the two-particle
continuum threshold !tp(q). The factor of 2, which arises
naturally, is due to the two fermions comprising a bosonic
molecule. Naturally, the presence of this term in the
uctuation-modied number equation is dependent upon
the existence of such a pole, that is, a molecular bound
state. These bound states correspond to the Feshbach
molecules in the presence of spin-orbit coupling and Zee-
man elds [7, 23].
Having extracted the pole contribution to Eq. (C4),
when it exists, the remaining integral over the branch
cut corresponds to scattering state fermions,
Nsc=TX
qZ1
!tp(q)d!
nB(!)@(q;!)
@ @(q;0)
@
;
(C6)
whose energy is larger than the minimum energy !tp(q)
of two free fermions. Thus, when bound states are
present, we arrive at the modied number equation,
N=N0+Nsc+Nb (C7)
whereN0is the number of free fermions obtained from
the saddle-point analysis in Eq. (B8), and NbandNsc
are the bound state and scattering contributions given in
Eqs. (C5) and (C6), respectively. These general results
are particularized to the equal Rashba-Dresselhaus case
in Sec. III B of this paper.
The number of unbound states Nuis then easily seen
to beNu=N0+Nsc, that is, the sum of the free-fermion
(N0) and scattering ( Nsc) contributions. Naturally, the
number of unbound states is also equal to the total num-
ber of states, N, minus the number of bound states, Nb,
that is,Nu=N Nb.
Appendix D: Derivation of Ginzburg-Landau
coecients for general Rashba-Dresselhaus
spin-orbit coupling
Next, we derive explicit expressions for the coe-
cients of the time-dependent Ginzburg-Landau theory
valid near the critical temperature of the super
uid. We
start from the
uctuation Lagrangian,
LF(r;) = T
2VTr ln( I+G0G 1
F) + ( r;) +j(r;)j2
g;
(D1)
in a volume V, and take the traces over both discrete
and continuous indices. Notice that the term ( r;) =
[0(r;) +
0(r;)]=gin the
uctuation Lagrangian
cancels out the linear terms in andwhen the log-
arithm is expanded, due to the saddle-point condition.Since the expansion is performed near Tc, we take the
saddle-point order parameter 0!0 and redene the
uctuation eld as (r;) = ( r;) to obtain
LF(r;) =jj2
g T
2VTr ln( I+G0[0]G 1
F[]):(D2)
Notice that the arguments in G0[0] and G 1
F[] represent
the values of 0= 0 and= , respectively.
We expand the logarithm to sixth order in to obtain
LF(r;) =jj2
g+T
2VTr1
2(G0G 1
F)2+1
4(G0G 1
F)4
+1
6(G0G 1
F)6+:::
; (D3)
where the higher-order odd (cubic and quintic) terms in
the order-parameter amplitudes expansion can be shown
to vanish due to conservation laws and energy or momen-
tum considerations.
The traces can be evaluated explicitly by using the
momentum-space inverse single-particle Green's function
G 1
0(k;k0) =A 1(k) 0
0
A 1( k)T
kk0;(D4)
derived from Eq. (A6). Here, we use the shorthand no-
tationk(i!;k), where!n= 2nT are bosonic Mat-
subara frequencies and dene the 2 2 matrix,
A 1(k) =
i!n eK"(k)h
?(k)
h?(k)i!n eK#(k)
; (D5)
whereeK"=k hz,eK#=k+hz, withk=k2=2m
the kinetic energy relative to the chemical potential, hz
the external Zeeman eld, and h?(k) =hx(k) +ihy(k)
the spin-orbit eld. We also dene the
uctuation con-
tribution to the inverse Green's function,
G 1
F(k;k0) =0 iyk k0
iyy
k0 k0
; (D6)
whereyis the second Pauli matrix in isospin (internal
state) space and
k=
VZ
0dZ
d3rei(kr !)(r) (D7)
is the Fourier transform of ( r), withr(r;), and also
has dimensions of energy. Recall that we set ~=kB= 1,
such that energy, frequency and temperature have the
same units.
Inversion of Eq. (D4) yields
G0(k;k0) =A(k) 0
0 [A( k)]T
kk0; (D8)
where the matrix A(k) is
A(k) =1
det[A 1(k)]
i!n eK#(k) h
?(k)
h?(k)i!n eK"k)
:
(D9)12
with det[ A 1(k)] =Q2
j=1[i!n Ej(k)] and where the
independent-particle eigenvalues Ej(k) are two of the
poles of G0(k;k). These poles are exactly the gen-
eral eigenvalues described in Eqs. (B2) in the limit of
0!0. Note that setting 0= 0 in the general eigen-
value expressions yields E1;2(k) =jjkjhkj. The other
set of poles of G0(k;k) corresponds to the eigenvalues
E3;4(k) = E2;1(k) found from det
A 1( k)T= 0.
Using Eq. (D3) to write the
uctuation action as SF=R
0dR
d3rLF(r;);results in
SF=VX
qjqj2
(q)+V
2X
q1;q2;q3b1;2;31
23
1 2+3
+V
3X
q1q5f151
23
45
1 2+3 4+5;(D10)
where summation over q(iqn;q) indicates sums over
both the bosonic Matsubara frequencies qn= 2nT and
momentum q. Here, we used the shorthand notation
jqjto represent the labels of qjor
qj.
The quadratic order appearing in Eq. (D10) arises
from the termsj(r;)j2=gand (T=2V)Tr(G0G 1
F)2=2
in Eq. (D3), and is directly related to the pair propaga-
torpair(q) =V 1(q), with
1(q) =1
g T
2VX
kTr
A(k)A 1(q k)
det[A 1(q k)];(D11)
where we use the identity yAy= det( A)(AT) 1:
The fourth-order contribution arises from1
4(G0G 1
F)4and leads to
b(q1;q2;q3) =T
2VX
kTr
A(k)A 1(q1 k)A(k q1+q2)A 1(q1 q2+q3 k)
det [A 1(q1 k)] det [ A 1(q1 q2+q3 k)]; (D12)
while the sixth order contribution emergences from1
6(G0G 1
F)6, giving
f(q1;;q5) =T
2VX
kdet [A(q1 k)] det [ A(q1 q2+q3 k)] det [ A(q1 q2+q3 q4+q5 k)]
Tr
A(k)A 1(q1 k)A(k q1+q2)A 1(q1 q2+q3 k)
A(k q1+q2 q3+q4)A 1(q1 q2+q3 q4+q5 k)
: (D13)
Evaluating the expressions given in Eqs. (D11)
through (D13) requires us to perform summations over
Matsubara frequencies of the type
TX
i!n1
i!nE(k)=(
n(k) if \+"
1 n(k) if \ ";(D14)
wheren(k) = 1=
eE(k)+ 1
is the Fermi function. For
the quadratic term, we obtain the result
1(q;iqn) = m
4as+1
2VX
k1
"k
+2X
i;j=1ij(k;q)Wij(k;q;iqn)
;(D15)
where the functions in the last term are
Wij(k;q;iqn) =1 ni(k) nj(k+q)
iqn Ei(k) Ej(k+q); (D16)corresponding to the contribution of bubble diagrams to
the pair susceptibility. The coherence factors are
11(k;q) =jukuk+q vkv
k+qj2; (D17)
12(k;q) =jukvk+q+uk+qvkj2; (D18)
with11(k;q) =22(k;q) and12(k;q) =21(k;q);
where the quasiparticle amplitudes are
uk=s
1
2
1 +hz
hk
; (D19)
vk=eiks
1
2
1 hz
hk
: (D20)
The anglekis the phase associated with the spin-orbit
eldh?(k) =jh?(k)jeik;and we replaced the interac-
tion parameter gby thes-wave scattering length asvia13
Eq. (B7), recalling that "k=k2=2m. The phase and
modulus of h?(k) are
k= arctanky
kx
; (D21)
jh?(k)j=jj
mq
k2x+k2y; (D22)
and the total eective eld is
hk=p
h2z+jh?(k)j2: (D23)Since we are interested only in the long-wavelength
and low-frequency regime, we perform an analytic con-
tinuation to real frequencies iqn=!+iafter calculat-
ing the Matsubara sums for all coecients appearing in
Eq. (D10) and perform a small momentum qand low-
frequency!expansion resulting in the Ginzburg-Landau
action,
SF=SGL=VX
q
a+X
`c`q2
`
2m d0!!
jqj2+V
2X
q1;q2;q3b(q1;q2;q3)q1
q2q3
q1 q2+q3
+V
3X
q1q5f(q1;q2;q3;q4;q5)q1
q2q3
q4q5
q1 q2+q3 q4+q5: (D24)
Here, the label `appearing explicitly in the termP
`c`q2
`=(2m) represents the spatial directions fx;y;zg,
while theqj's in the sums correspond to ( qj;!j) and
the summationsP
qjrepresent integrals VR
d!jR
d3qj,
wherejlabels a fermion pair and can take values in the
setf1;2;3;4;5g. In the expression above, we used the
result
1(q;!) =a+X
`c`q2
`
2m d0!+ (D25)
for the analytically continued expression of 1(q;iqn)
appearing in Eq. (D15). To write the coecients above
in a more compact notation, we dene
Xi=Xi(k) = tanh [Ei(k)=2]; (D26)
Yi=Yi(k) = sech2[Ei(k)=2]: (D27)
The frequency- and momentum-independent coecient
is
a= m
4as+1
VX
k1
2"k X1
4E1+X2
4E2
h2
z
khkX1
4E1 X2
4E2
;(D28)whereE1=E1(k) andE2=E2(k). The coecient
d0=dR+idImultiplying the linear term in frequency
has a real component given by
dR=1
2VPX
k2X
i;j=1ij(k;0)1 ni(k) nj(k)
[Ei(k) +Ej(k)]2:(D29)
Using the explicit forms of the coherence factors ukand
vkthat dene ij(k;q=0), the above expression can be
rewritten as
dR=1
2VPX
k
1 +h2
z
2
kX1
4E2
1+X2
4E2
2
+2h2
z
khkX1
4E2
1 X2
4E2
2
;(D30)
which denes the time scale for temporal oscillations of
the order parameter. Here, the symbol Pdenotes the
principal value, and the coecient dRis obtained from
Re
1(q=0;!+i)
= m
4as+1
2VX
k2
41
"k+P2X
i;j=1ij(k;q=0)1 ni(k) nj(k)
! Ei(k) Ej(k)3
5: (D31)
The imaginary component of the coecient dhas the form
dI=
2VX
k2X
i;j=1ij(k;0) [1 ni(k) nj(k)]0(Ei(k) +Ej(k)); (D32)14
where the derivative of the delta function is 0() =@(x+)=@xjx=0:Using again the expressions of the coherence
factorsukandvkleads to
dI=
2VX
k
(X1+X2)0(2k) +jh?j2
h2
k
X10(2E1) +X20(2E2) (X1+X2)0(2k)
; (D33)
which determines the lifetime of fermion pairs. This result originates from
Im
1(q=0;!+i)
=
2VX
k2X
i;j=1ij(k;q=0) [1 ni(k) nj(k)](! Ei(k) Ej(k)); (D34)
which immediately reveals that below the two-particle
threshold!tp(q=0) = minfi;j;kg[Ei(k) +Ej(k)] at
center-of-mass momentum q=0, the lifetime of the pairs
is innitely long due to the emergence of stable two-body
bound states. Note that collisions between bound states
are not yet included.
The expressions for the c`coecients appearing in
Eq. (D25) are quite long and complex. Since these coef-
cients are responsible for the mass renormalization and
anisotropy within the Ginzburg-Landau theory, we out-
line below their derivation in detail. These coecients
can be obtained from the last term in Eq. (D15), which
we dene as
F(q) =1
2VX
k2X
i;j=1ij(k;q)Wij(k;q;iqn= 0):(D35)
The relation between c`and the function F(q) dened
above is
c`=m@2F(q)
@q2
`
q=0: (D36)
A more explicit form of c`is obtained by analyzing the
symmetry properties of F(q) under inversion and re
ec-
tion symmetries. To make these properties clear, we
rewrite the summand in Eq. (D35) by making use of the
transformation k!k q=2. This procedure leads to the
symmetric form,
F(q) =1
2VX
k2X
i;j=1eij(k ;k+)fWij[Ei(k );Ej(k+)]:
(D37)
Here, k+=k+q=2 and k =k q=2 are new momentum
labels, and
e11(k ;k+) =juk uk+ vk v
k+j2; (D38)
e12(k ;k+) =juk vk+ vk uk+j2(D39)
are coherence factors, with e11(k ;k+) =e22(k ;k+)
ande12(k ;k+) =e21(k ;k+) The functions ukand
vkare dened in Eqs. (D19) and (D20). It is now very
easy to show that eij(k ;k+) =eij(k+;k ), that is,
eij(k ;k+) is an even function of q, since taking q! q
leads to k !k+andk+!k leavingeijinvariant.It is also clear, from its denition, that eijis symmetric
in the band indices fi;jg. Furthermore, the function
fWij[Ei(k );Ej(k+)] =Nij
Dij; (D40)
dened above, is the ratio between the numerator,
Nij= tanh [Ei(k )=2] + tanh [Ej(k+)=2];(D41)
representing the Fermi occupations and the denominator,
Dij= 2 [Ei(k ) +Ej(k+)]; (D42)
representing the sum of the quasi-particle excitation
energies. To elliminate the Fermi distributions ni(k)
in the numerator, we used the relation 1 2ni(k) =
tanh [Ei(k )=2]. Notice that fWij[Ei(k );Ej(k+)] is
not generally symmetric under inversion q! q,
that is, under the transformation k !k+and
k+!k . This means that fWij[Ei(k );Ej(k+)]6=
fWij[Ei(k+);Ej(k )], unless when i=j, where
it is trivially an even function of q. However,
fWij[Ei(k );Ej(k+)] is always symmetric under simul-
taneous momentum inversion ( q! q) and band index
exchange, that is,
fWij[Ei(k );Ej(k+)] =fWji[Ej(k+);Ei(k )] (D43)
for anyfi;jg. This property will be used later to write a
nal expression for c`. Next, we write
@2F(q)
@q2
`
q=0=1
2VX
k2X
i;j=1Fij; (D44)
where the function inside the summation is
Fij="
@2eij
@q2
`fWij+ij@2fWij
@q2
`#
q=0: (D45)
Notice the absence of terms containing the product of
the rst-order derivatives of eijandfWij. These terms
vanish due to parity since eijis an even function of q,
leading to [@eij=@q`]q=0= 0. The last expression can be15
further developed upon summation over the band indices,
leading to
@2F(q)
@q2
`
q=0=A+B: (D46)
The rst contribution is given by
A=1
2VX
k@2e11
@q2
`fWdi+@2e12
@q2
`fWod
q=0;(D47)
and contains the second derivatives of eijand the sym-
metric terms
fWdi=
fW11+fW22
; (D48)
fWod=
fW12+fW21
; (D49)The second contribution is given by
B=1
2VX
k"
e11@2fWdi
@q2
`+e12@2fWod
@q2
`#
q=0:(D50)
Next, we explicitly write eij,fWijand their second
derivatives with respect to q`atq=0. We start with
h
fWiji
q=0=Xi+Xj
2 [Ei+Ej](D51)
and for the second derivative, we write
"
@2fWij
@q2
`#
q=0=1
Dij@2Nij
@q2
`
q=0 "
2
D2
ij@Dij
@q`@Nij
@q`#
q=0+"
2Nij
D3
ij@Dij
@q`2#
q=0 "
Nij
D2
ij@2Dij
@q2
`#
q=0:
Each one of the four terms in the above expression is
evaluated at q=0and can be written in terms of specic
expressions that are given below. The numerator is
[Nij]q=0=Xi+Xj; (D52)
the rst derivative of Nijis
@Nij
@q`
q=0=Y2
j
4T@Ej
@k` Y2
i
4T@Ei
@k`; (D53)
and the second derivative of Nijis
@2Nij
@q2
`
q=0= XjY2
j
8T2@Ej
@k`2
+Yi
8T@2Ei
@k2
`
XiY2
i
8T2@Ei
@k`2
+Y2
i
8T@2Ei
@k2
`:(D54)
The denominator Dijand its rst derivative are
[Dij]q=0= 2(Ei+Ej); (D55)
@Dij
@q`
q=0=@Ej
@k` @Ei
@k`; (D56)
while the second derivative of Dijis
@2Dij
@q2
`
q=0=1
2@2Ei
@k2
`+@2Ej
@k2
`
: (D57)
When the order parameter is zero, that is, j0j= 0, the
energiesE1(k) andE2(k) become
E1(k) =jkj+hk (D58)
E2(k) =jkj hk: (D59)The rst derivatives of these energies are
@E1(k)
@k`=S1(k)k`
m+@hk
@k`; (D60)
@E2(k)
@k`=S2(k)k`
m @hk
@k`; (D61)
with the functions S1(k) = sgn [jkj+hk] sgn [k] and
S2(k) = sgn [jkj hk] sgn [k]:The derivative of the ef-
fective Zeeman eld is
@hk
@k`=1
hk2
m2(kx`x+ky`y): (D62)
The second derivatives of the energies are
@2E1(k)
@k2
`=S1(k)
m+@2hk
@k2
`(D63)
@2E2(k)
@k2
`=S2(k)
m @2hk
@k2
`; (D64)
where the second derivative of the eective eld is
@2hk
@k2
`=1
hk2
m2
(`x+`y) 1
h2
k2
m2
k2
x`x+2k2
y`y
:
(D65)
Since the diagonal elements fWiiare even functions of
qand so areNiiandDii, their expressions are simpler
than in the general case discussed above, because the rst
order derivatives of NiiandDiivanish. The surviving
terms involve only the second derivatives of NiiandDii
leading to the expression
"
@2fWii
@q2
`#
q=0=1
Dii@2Nii
@q2
`
q=0 Nii
D2
ii@2Dii
@q2
`
q=0:
(D66)16
Here, the numerator and denominator functions are
[Nii]q=0= 2Xiand [Dii]q=0= 4Ei; (D67)
while their second derivatives are
@2Nii
@q2
`
q=0= XiY2
i
4T2@Ei
@k`2
+Y2
i
4T@2Ei
@k2
`;(D68)
@2Dii
@q2
`
q=0=@2Ei
@k2
`: (D69)
The next step in obtaining the c`coecients is to an-
alyze the functions eijand their second derivatives. We
begin by writing e11atq=0:
[e11]q=0=u2
k jvkj22=h2
z
h2
k: (D70)
To investigate the second derivative of e11, we write
e11=
11
11; (D71)
where the complex function is given by
11=uk uk+ vk vk+: (D72)
In this case, we write the rst derivative of e11as
@e11
@q`=@
11
@q`
11+
11@
11
@q`(D73)
and the second derivative as
@2e11
@q2
`=@2
11
@q2
`
11+ 2@
11
@q`@
11
@q`+
11@2
11
@q2
`:(D74)
To explore the symmetry with respect to q, we express
11in terms of its odd and even components via the re-
lation
11=
11;e+
11;o, where the even component
11;e= [
11(q) +
11( q)]=2 is
11;e=uk uk+ jvk jjvk+jcos
k k+
(D75)
and the odd component
11;o= [
11(q)
11( q)]=2 is
11;o=ijvk+jjvk jsin
k+ k
: (D76)
Expressed via the even
11;eand odd
11;ocomponents,
the second derivative in Eq. (D74) is
@2e11
@q2
`=@2
11;e
@q2
`
11:e+ 2@
11;o
@q`@
11;o
@q`+
11;e@2
11;e
@q2
`:
(D77)
Notice that the even component is purely real, that is,
11;e=
11;e, and that the odd component is purely imag-
inary,
11;o=
11;o. Use of this property leads to
@2e11
@q2
`= 2
11;e@2
11;e
@q2
` 2@
11;o
@q`2
: (D78)The contribution from the even term
11;eis
[
11;e]q=0=u2
k jvkj2=hz
hk; (D79)
and from its second derivative is
@2
11;e
@q2
`
q=0=1
2@jvkj
@k`2
1
2jvkj@2jvkj
@k2
`+jvkj2@k
@k`2
;
(D80)
while the contribution from the odd term
11;ois
@
11;o
@q`
q=0=ijvkj2@k
@k`: (D81)
Now, we turn our attention to e12and its second
derivative. From Eq. (D39), we notice that
12is ex-
plicitly odd in qbecause
12(q) =
12( q), since the
operation q! qtakes k !k+and vice versa, leading
to
[e12]q=0= 0: (D82)
To calculate the second derivative of e12, we write
e12=
12
12; (D83)
where the complex function
12=uk vk+ vk uk+: (D84)
We relate @2e12=@q2
`to
12and its rst and second
derivatives via
@2e12
@q2
`=@2
12
@q2
`
12+ 2@
12
@q`@
12
@q`+
12@2
12
@q2
`:(D85)
Given that [
12]q=0= 0 and [
12]q=0= 0, the expression
above simplies to
@2e12
@q2
`
q=0= 2@
12
@q`@
12
@q`
q=0= [`(q)]2;(D86)
where we used the expressions
@
12
@q`
q=0=eik`(k) (D87)
for the derivatives of
12atq=0with the function
`(k) =uk@jvkj
@k` jvkj@uk
@k`+ukjvkj@k
@k`: (D88)
The last information needed is the derivatives of uk,
jvkj, andk, which are given by
@uk
@k`= 1
2hz
h3
k2
m2(kx`x+ky`y)
(1 +hz=hk)1=2; (D89)
@jvkj
@k`=1
2hz
h3
k2
m2(kx`x+ky`y)
(1 hz=hk)1=2; (D90)
@k
@k`=(kx`y ky`x)
k2x+2k2y: (D91)17
The long steps discussed above complete the derivation of
all the functions needed to compute the c`coecients for
an arbitrary spin-orbit coupling, expressed as a general
linear combination of Rashba and Dresselhaus terms.
As announced earlier, the calculation of c`, dened in
Eq. (D36), is indeed very long and requires the use of
all the expressions given from Eq. (D37) to Eq. (D91).
Despite this complexity, that are a few important com-
ments about the symmetries of the c`coecients that
are worth mentioning. Given that c`determines the
mass anisotropies in the Ginzburg-Landau (GL) theory,
we discuss next the anisotropies of c`as a function of
the spin-orbit coupling parameters and. First, in
the limit of zero spin-orbit coupling, where andare
equal to zero, all the c`coecients are identical re
ect-
ing the isotropy of the system, that is, cx=cy=cz
and reduce to previously known results [52]. In this
case, the GL eective masses m`=mdR=c`are isotropic:
mx=my=mz. Second, in the limit of 6= 0 and=1, the spin-orbit coupling has the same strength
along thexandydirections, and thus for the Rashba
(= 1) or Dresselhaus ( = 1) cases, the coecients
obey the relation cx=cy6=cz. This leads to eective
massesmx=my6=mz. Third, in the limit 6= 0, but
= 0, corresponding to the ERD case, the coecients
have the symmetry cx6=cy=cz. Now the eective
masses obey the relation mx6=my=mz. Finally, in
the case where 6= 0, and 06=jj<1, all thec`coe-
cients are dierent, that is, cx6=cy6=cz. Therefore, the
eective masses are also dierent in all three directions:
mx6=my6=mz.
Following an analogous procedure, we analyze the co-
ecientsb(q1;q2;q3), ande(q1;q2;q3;q4;q5) with allqi=
(0;0), and dene
Zij=Xi+EiYj=2: (D92)
Using the notation b(0;0;0) =b(0), we obtain
b(0) =1
8VX
k
1 +h4
z
2
kh2
kZ11
E3
1+Z22
E3
2
+2h2
z
khkZ11
E3
1 Z22
E3
2
+h4
z
3
kh3
kX1
E1 X2
E2
; (D93)
which is a measure of the local interaction between two pairing elds. Using the notation f(0;0;0;0;0) =f(0), we
obtain
f(0) =3
32VX
k
1 +3h4
z
2
kh2
kZ11
E5
1+Z22
E5
2
h2
z
khk
3 +h4
z
2
kh2
kZ11
E5
1 Z22
E5
2
h6
z
4
kh4
kZ11
E3
1+Z22
E3
2
h4
z
3
kh3
kZ11
E3
1 Z22
E3
2
+2
6X1Y1
E3
1+X2Y2
E3
2
+2h2
z
6khkX1Y1
E3
1 X2Y2
E3
2
h6
z
5
kh5
kX1
E1 X2
E2
; (D94)
which is a measure of the local interaction between three pairing elds. It is important to mention that in the absence
of spin-orbit and Zeeman elds, the Ginzburg-Landau coecients obtained above reduce to those reported in the
literature [52].
As we proceed to explicitly write the Ginzburg-Landau
action and Lagrangian density, we emphasize that in con-
trast to the standard crossover that one observes in the
absence of an external Zeeman eld [52], for xed hz6= 0
it is possible for the system to undergo a rst-order phase
transition with increasing 1 =kFas. The same applies for
xed 1=kFaswith increasing hz. Thus, while an expan-
sion ofSFto quartic order is sucient when no Zeeman
elds are present, when Zeeman elds are turned on, the
fourth-order coecient b(0) =bmay become negative.
Such a situation requires the analysis of the sixth-order
coecientf(0) =fto describe this rst-order transition
correctly and to stabilize the theory since f >0.
The Ginzburg-Landau action in Euclidean space can
be written asSGL=R
dtR
d3rLGL(r);wherer(r;t).Here, the Lagrangian density is
LGL(r)=aj(r)j2+b
2j(r)j4+f
3j(r)j6
+X
`c`jr`(r)j2
2m id0(r)@(r)
@t;(D95)
where`=fx;y;zg,b=b(0) andf=f(0). A variation
ofSGLwith respect to (r) viaSGL=(r) = 0 yields
the time-dependent Ginzburg-Landau (TDGL) equation,
id0@
@t X
`c`r2
`
2m+bjj2+fjj4+a!
(r) = 0
(D96)
with cubic and quintic terms, where = ( r) are de-
pendent on space and time. This equation describes the
spatio-temporal behavior of the order parameter ( r;t)
in the long-wavelength and long-time regime.18
In the static homogeneous case with b>0, Eq. (D96)
reduces to either the trivial (normal-state) solution =
0 whena > 0 or to the nontrivial (super
uid state)
jj=p
jaj=b, whena < 0. The coecient dprovides
the timescale of the TDGL equation, and thereby deter-
mines the lifetime associated with the pairing eld ( r).
This can be seen directly by again considering the ho-
mogeneous case to linear order in ( r), in which case
the TDGL equation has the solution ( t)(0)eiat=d 0:
This last expression can be rewritten more explicitly
as (t)(0)e i!0te t=0;where!0=jajdR=jd0j2is
the oscillation frequency of the pairing eld, and 0=
jd0j2=(jajdI) is the lifetime of the pairs, where both dR
anddIare positive denite, that is, dR>0 anddI>0.
In the BEC regime, where stable two-body bound
states exist, the imaginary part of d0vanishes (dI= 0),
and the lifetime time of the pairs is innitely long. In
this case,d0=dRand we can dene the eective bosonic
wave function =pdR to recast Eq. (D96) in the form
of the Gross-Pitaevskii equation,
i@
@t X
`r2
`
2M`+U2j j2+U3j j4 B!
(r) = 0;
(D97)with cubic and quintic nonlinearities, where = ( r),
to describe a dilute Bose gas. Here, B= a=dRis
the bosonic chemical potential, M`=m(dR=c`) are the
anisotropic masses of the bosons, and U2=b=d2
Rand
U3=f=d3
Rrepresent contact interactions of two and
three bosons, respectively. In the Bose regime, the life-
timeof the composite boson is /1=dI!1 and the
interactions U2andU3are always repulsive, thus leading
to a system consisting of a dilute gas of stable bosons.
In this regime, the chemical potential of the bosons is
B2+Eb<0, whereEbis the two-body bound state
energy in the presence of spin-orbit coupling and Zeeman
elds obtained from the condition 1(q;E 2) = 0
discussed in the main text. Notice that when B!0 ,
in the absence of boson-boson interactions, the bosons
condense.
[1] Y-J. Lin, R. L. Compton, K. Jimin ez-Garc a, J. V. Porto,
and I. B. Spielman, Synthetic magnetic elds for ultra-
cold neutral atoms, Nature (London) 462, 628 (2009).
[2] Y-J. Lin, R. L. Compton, A. R. Perry, W. D. Phillips, J.
V. Porto, and I. B. Spielman, Bose-Einstein condensate
in a uniform light-induced vector potential, Phys. Rev.
Lett. 102, 130401 (2009).
[3] C. J. Kennedy, W. C. Burton, W. C. Chung and W.
Ketterle, Observation of Bose{Einstein condensation in
a strong synthetic magnetic eld, Nature Physics, 11,
859 (2015).
[4] Y.-J. Lin, K. Jim enez-Garc a and I. B. Spielman,
Spin{orbit-coupled Bose{Einstein condensates, Nature
471, 83 (2011).
[5] L. W. Cheuk, A. T. Sommer, Z. Hadzibabic, T. Yefsah,
W. S. Bakr, and M. W. Zwierlein, Spin-injection spec-
troscopy of a spin-orbit coupled Fermi gas, Phys. Rev.
Lett. 109, 095302 (2012).
[6] P. Wang, Z.-Q. Yu, Z, Fu, J, Miao, L, Huang, S. Chai, H.
Zhai, and J. Zhang, Spin-orbit coupled degenerate Fermi
gases, Phys. Rev. Lett. 109, 095301 (2012).
[7] R. A. Williams, M. C. Beeler, L. J. LeBlanc, K. Jim enez-
Garc a and I. B. Spielman, Raman-Induced Interactions
in a Single-Component Fermi Gas Near an s-Wave Fesh-
bach Resonance, Phys. Rev. Lett. 111, 095301 (2013).
[8] V. Galitski and I. B. Spielman, Spin{orbit coupling in
quantum gases, Nature 494, 49 (2013).
[9] Z. Fu, L. Huang, Z. Meng, P. Wang, L. Zhang, S. Zhang,
H. Zhai, P. Zhang, and J. Zhang, Production of Feshbach
molecules induced by spin-orbit coupling in Fermi gases,
Nat. Phys. 10, 110 (2014).
[10] M. Mancini, G. Pagano, G. Cappellini, L. Livi, M. Rider,J. Catani, C. Sias, P. Zoller, M. Inguscio, M. Dalmonte,
and L. Fallani, Observation of chiral edge states with
neutral fermions in synthetic Hall ribbons, Science 349,
1510 (2015).
[11] L. Huang, Z. Meng, P. Wang, P. Peng, S.-L. Zhang, L.
Chen, D. Li, Q. Zhou, and J. Zhang, Experimental real-
ization of two-dimensional synthetic spin{orbit coupling
in ultracold Fermi gases, Nature Physics 12, 540 (2016).
[12] Z. Wu, L.Zhang, W. Sun, X.-T. Xu, B.-Z. Wang, S.-C. Ji,
Y. Deng, S. Chen, X.-J. Liu, and J.-W. Pan, Realization
of two-dimensional spin-orbit coupling for Bose-Einstein
condensates Science 354, 83 (2016).
[13] J. I. Cirac and P. Zoller, Goals and opportunities in quan-
tum simulation, Nature Phys. 8, 264 (2012).
[14] U.-J. Wiese, Ultracold quantum gases and lattice sys-
tems: Quantum simulation of lattice gauge theories, An-
nalen der Physik 525, 777 (2013).
[15] E. Zohar, J. I. Cirac, and B. Reznik, Quantum simula-
tions of lattice gauge theories using ultracold atoms in
optical lattices, Rep. Prog. Phys. 79, 1 (2015).
[16] M. Dalmonte and S. Montangero, Lattice gauge theory
simulations in the quantum information era, Contemp.
Phys. 57, 388 (2016).
[17] M. Gong, S. Tewari, and C. Zhang, BCS-BEC crossover
and topological phase transition in 3D spin-orbit cou-
pled degenerate Fermi gases Phys. Rev. Lett. 107, 195303
(2011).
[18] Z.-Q. Yu and H. Zhai, Spin-orbit coupled Fermi gases
across a Feshbach resonance, Phys. Rev. Lett. 107,
195305 (2011).
[19] H. Hu, L. Jiang, X.-J. Liu, and H. Pu, Probing
anisotropic super
uidity in atomic Fermi gases with19
Rashba spin-orbit coupling, Phys. Rev. Lett. 107, 195304
(2011).
[20] L. Han and C. A. R. S a de Melo, Evolution from BCS to
BEC super
uidity in the presence of spin-orbit coupling,
Phys. Rev. A 85, 011606(R) (2012).
[21] K. Seo, L. Han, and C. A. R. S a de Melo, Emergence
of Majorana and Dirac particles in ultracold fermions
via tunable interactions, spin-orbit eects, and Zeeman
elds, Phys. Rev. Lett. 109, 105303 (2012).
[22] K. Seo, L. Han, and C. A. R. S a de Melo, Topological
phase transitions in ultracold Fermi super
uids: The evo-
lution from Bardeen-Cooper-Schrieer to Bose-Einstein-
condensate super
uids under articial spin-orbit elds,
Phys. Rev. A 85, 033601 (2012).
[23] D. M. Kurkcuoglu and C. A. R. S a de Melo, Formation of
Feshbach molecules in the presence of articial spin-orbit
coupling and Zeeman elds, Phys. Rev. A 93, 023611
(2016).
[24] N. Goldman, I. Satija, P. Nikolic, A. Bermudez, M. A.
Martin-Delgado, M. Lewenstein, and I. B. Spielman, Re-
alistic time-reversal invariant topological Insulators with
neutral atoms, Phys. Rev. Lett. 105, 255302 (2010).
[25] I. B. Spielman, private communication.
[26] G. Juzeli unas, J. Ruseckas, and J. Dalibard, Generalized
Rashba-Dresselhaus spin-orbit coupling for cold atoms,
Phys. Rev. A 81, 053403 (2010).
[27] D. L. Campbell, G. Juzeli unas, and I. Spielman, Realistic
Rashba and Dresselhaus spin-orbit coupling for neutral
atoms Phys. Rev. A 84, 025602 (2011).
[28] G. Dresselhaus, Spin-orbit coupling eects in zinc blende
structures Phys. Rev. 100, 580 (1955).
[29] E. I. Rashba, Properties of semiconductors with an ex-
tremum loop: I. Cyclotron and combinational resonance
in a magnetic eld perpendicular to the plane of the loop,
Sov. Phys. Solid State 2, 1224 (1960)
[30] C. A. R. S a de Melo, When fermions become bosons:
Pairing in ultracold gases, Physics Today vol. 61issue
10, page 45 (2008).
[31] T. Ozawa and G. Baym, Population imbalance and pair-
ing in the BCS-BEC crossover of three-component ultra-
cold fermions, Phys. Rev. A 82, 063615 (2010).
[32] P. D. Powell, From Quarks to Cold Atoms: The Phases of
Strongly-Interacting Systems, Ph.D. Thesis, University
of Illinois at Urbana-Champaign (2013).
[33] D. M. Kurkcuoglu and C. A. R. S a de Melo, Color super-
uidity of neutral ultracold fermions in the presence of
color-
ip and color-orbit elds, Phys. Rev. A 97, 023632
(2018)
[34] Daisuke Yamamoto, I. B. Spielman, and C. A. R. S a de
Melo, Quantum phases of two-component bosons with
spin-orbit coupling in optical lattices, Phys. Rev. A 96,
061603(R) (2017).
[35] S. L. Bromley, S. Kolkowitz, T. Bothwell, D. Kedar, A.
Safavi-Naini, M. L. Wall, C. Salomon, A. M. Rey, and
J. Ye, Dynamics of interacting fermions under spin{orbit
coupling in an optical lattice clock, Nature Physics 14,
399 (2018).
[36] Yu Yi-Xiang, Fadi Sun, and Jinwu Ye, Class of topo-
logical phase transitions of Rashba spin-orbit coupled
fermions on a square lattice, Phys. Rev. B 98, 174506
(2018).
[37] M. Mamaev, R. Blatt, J Ye, and A. M. Rey, Cluster
state generation with spin-orbit coupled fermionic atoms
in optical lattices, Phys. Rev. Lett. 122, 160402 (2019).[38] Baihua Gong, Shuai Li, Xin-Hui Zhang, Bo Liu, and Wei
Yi, Bloch bound state of spin-orbit-coupled fermions in
an optical lattice, Phys. Rev. A 99, 012703 (2019).
[39] Man Hon Yau and C. A. R. S a de Melo Chern-number
spectrum of ultracold fermions in optical lattices tuned
independently via articial magnetic, Zeeman, and spin-
orbit elds, Phys. Rev. A 99, 043625 (2019)
[40] Wei Jia, Zhi-Hao Huang, Xian Wei, Qing Zhao, and
Xiong-Jun Liu, Topological super
uids for spin-orbit cou-
pled ultracold Fermi gases, Phys. Rev. B 99, 094520
(2019).
[41] Majid Kheirkhah, Zhongbo Yan, Yuki Nagai, and Frank
Marsiglio, First- and second-order topological supercon-
ductivity and temperature-driven topological phase tran-
sitions in the extended hubbard model with spin-orbit
coupling, Phys. Rev. Lett. 125, 017001 (2020).
[42] Urs Gebert, Bernhard Irsigler, and Walter Hofstetter,
Local Chern marker of smoothly conned Hofstadter
fermions, Phys. Rev. A 101, 063606 (2020).
[43] Irakli Titvinidze, Julian Legendre, Maarten Grothus,
Bernhard Irsigler, Karyn Le Hur, and Walter Hofstet-
ter, Spin-orbit coupling in the kagome lattice with
ux
and time-reversal symmetry, Phys. Rev. B 103, 195105
(2021).
[44] R. A. Williams, L. J. LeBlanc, K. Jim enez-Garc a, M. C.
Beeler, A. R. Perry, W. D. Phillips, and I. B. Spielman,
Synthetic partial waves in ultracold atomic collisions, Sci-
ence335, 314 (2012).
[45] Z. Meng, L. Huang, P. Peng, D. Li, L. Chen, Y. Xu,
C. Zhang, P. Wang, and J. Zhang Experimental obser-
vation of a topological band gap opening in ultracold
Fermi gases with two-dimensional spin-orbit coupling,
Phys. Rev. Lett. 117, 235304 (2016).
[46] L. Huang, Z. Meng, P. Wang, P. Peng, S.-L. Zhang, L.
Chen, D. Li, Q. Zhou, and J. Zhang, Experimental real-
ization of two-dimensional synthetic spin{orbit coupling
in ultracold Fermi gases, Nature Physics 12, 540 (2016).
[47] A. Vald es-Curiel, D. Trypogeorgos, Q.-Y. Liang, R. P.
Anderson, and I. B. Spielman, Topological features with-
out a lattice in Rashba spin-orbit coupled atoms, Nature
Communications 12, 593 (2021).
[48] Ronen M. Kroeze, Yudan Guo, and Benjamin L. Lev,
Dynamical spin-orbit coupling of a quantum gas Phys.
Rev. Lett. 123, 160404 (2019).
[49] Zong-Yao Wang, Xiang-Can Cheng, Bao-Zong Wang,
Jin-Yi Zhang, Yue-Hui Lu, Chang-Rui Yi, Sen Niu, You-
jin Deng, Xiong-Jun Liu, Shuai Chen, and Jian-Wei Pan,
Realization of an ideal Weyl semimetal band in a quan-
tum gas with 3D spin-orbit coupling, Science 372, 271
(2021).
[50] A. J. Leggett in Modern Trends in the Theory of Con-
densed Matter , pp. 13{27, edited by A. Pekalski and R.
Przystawa, Springer-Verlag, Berlin (1980).
[51] J. R. Engelbrecht, M. Randeria, and C. A. R. S a de Melo,
BCS to Bose crossover: Broken-symmetry state, Phys.
Rev. B 55, 15153 (1997).
[52] C. A. R. S a de Melo, M. Randeria, and J. Engelbrecht,
Crossover from BCS to Bose superconductivity: Transi-
tion temperature and time-dependent Ginzburg-Landau
theory, Phys. Rev. Lett. 71, 3202 (1993).
[53] P. D. Powell, G. Baym, C. A. R. S a de Melo, Super
uid
transition temperature of spin-orbit and Rabi coupled
fermions with tunable interactions, arXiv:1709.07042v1
(2017).20
[54] S. Gopalakrishnan, A. Lamacraft, and P. M. Gold-
bart, Universal phase structure of dilute Bose gases with
Rashba spin-orbit coupling, Phys. Rev. A 84, 061604
(2011).
[55] T. Ozawa, Topics in Multi-component Ultracold Gases
and Gauge Fields, Ph.D. Thesis, University of Illinois at
Urbana-Champaign (2012).
[56] P. Nozi eres and S. Schmitt-Rink, Bose condensation in
an attractive fermion gas: From weak to strong coupling
superconductivity, J. Low Temp. Phys. 59, 195 (1985).
[57] Z. Yu and G. Baym, Spin correlation functions in ul-
tracold paired atomic-fermion systems: Sum rules, self-
consistent approximations, and mean elds, Phys. Rev.
A73, 063601 (2006).
[58] Z. Yu, G. Baym, and C. J. Pethick, Calculating energy
shifts in terms of phase shifts, J. Phys. B: At. Mol. Opt.
Phys. 44, 195207 (2011).
[59] P. Fulde and R. A. Ferrell, Superconductivity in a strong
spin-exchange eld, Phys. Rev. 135, A550 (1964).
[60] A. I. Larkin and Y. N. Ovchinnikov, Nonuniform state of
superconductors, Sov. Phys. JETP 20, 762 (1965).
[61] A. M. Clogston, Upper limit for the critical eld in hard
superconductors, Phys. Rev. Lett. 9, 266 (1962).
[62] G. Sarma, On the in
uence of a uniform exchange eld
acting on the spins of the conduction electrons in a su-
perconductor, J. Phys. Chem. Sol. 24, 1029 (1963).
[63] G. Baym, J.-P. Blaizot, M. Holzmann, F. Lalo e and D.
Vautherin, The transition temperature of the dilute in-
teracting Bose gas, Phys. Rev. Lett. 83, 1703 (1999).
[64] M. Iskin and C. A. R. S a de Melo, Fermi-Fermi mixtures
in the strong-attraction limit, Phys. Rev. A 77, 013625
(2008).
[65] D. S. Petrov, C. Salomon, and G. V. Shlyapnikov, Di-
atomic molecules in ultracold Fermi gases: Novel com-
posite bosons, J. Phys. B 38, S645 (2005).
[66] J. P. Vyasanakere, S. Zhang, and V. B. Shenoy, BCS-
BEC crossover induced by a synthetic non-Abelian gauge
eld, Phys, Rev. B 84, 014512 (2011)
[67] X.-J. Feng and L. Yin, Phase diagram of a spin-orbit
coupled dipolar Fermi gas at T= 0 K, Chin. Phys. Lett.
37, 020301 (2020).[68] L. Dell'Anna, G. Mazzarella, and L. Salasnich, Conden-
sate fraction of a resonant Fermi gas with spin-orbit cou-
pling in three and two dimensions, Physical Review A
84, 033633 (2011).
[69] L. Dell'Anna, G. Mazzarella, and L. Salasnich, Tuning
Rashba and Dresselhaus spin-orbit couplings: Eects on
singlet and triplet condensation with Fermi atoms, Phys-
ical Review A 86, 053632 (2012).
[70] S. Tewari, T. D, Stanescu, J. D. Sau and S. Das
Sarma, Topologically non-trivial superconductivity in
spin{orbit-coupled systems: Bulk phases and quantum
phase transitions, New Journal of Physics, 13, 065004
(2011).
[71] Li Han and C. A. R. S a de Melo, Ultra-cold fermions
in the
atland: Evolution from BCS to Bose super
uid-
ity in two-dimensions with spin-orbit and Zeeman elds,
arXiv:1206.4984v1 (2012)
[72] J. P. A. Devreese, J. Tempere, and C. A. R. S a de
Melo Eects of spin-orbit coupling on the Berezinskii-
Kosterlitz-Thouless transition and the vortex-antivortex
structure in two-dimensional Fermi gases, Phys. Rev.
Lett. 113, 165304 (2014).
[73] J. P. A. Devreese, J. Tempere, and C. A. R. S a de Melo
Quantum phase transitions and Berezinskii-Kosterlitz-
Thouless temperature in a two-dimensional spin-orbit-
coupled Fermi gas, Phys. Rev. A 92, 043618 (2015)
[74] J. P. Vyasanakere and V. B. Shenoy, Fluctuation theory
of Rashba Fermi gases: Gaussian and beyond, Phys. Rev.
B.92, 121111(R) (2015).
[75] T. Yamaguchi, D. Inotani, and Y. Ohashi, Rashbon
bound states associated with a spherical spin{orbit cou-
pling in an ultracold Fermi gas with an s-wave interac-
tion, J. Low Temp. Phys. 183, 161 (2016).
[76] B. M. Anderson, C.-T. Wu, R. Boyack, and K. Levin
Topological eects on transition temperatures and re-
sponse functions in three-dimensional Fermi super
uids,
Phys. Rev. B 92, 134523 (2015).
[77] L. Dell'Anna, and S. Grava, Critical temperature in the
BCS-BEC crossover with spin-orbit coupling, Condens.
Matter 6, 16 (2021). |
1110.6364v1.Artificial_spin_orbit_coupling_in_ultra_cold_Fermi_superfluids.pdf | arXiv:1110.6364v1 [cond-mat.quant-gas] 28 Oct 2011Artificial spin-orbit coupling in ultra-cold Fermi superflu ids
Kangjun Seo, Li Han and C. A. R. S´ a de Melo
School of Physics, Georgia Institute of Technology, Atlant a, Georgia 30332, USA
(Dated: November 23, 2018)
The control and understanding of interactions in many parti cle systems has been a major chal-
lenge in contemporary science, from atomic to condensed mat ter and astrophysics. One of the
most intriguing types of interactions is the so-called spin -orbit coupling - the coupling between the
spin (rotation) of a particle and its momentum (orbital moti on), which is omnipresent both in the
macroscopic and microscopic world. In astrophysics, the sp in-orbit coupling is responsible for the
synchronization of the rotation (spinning) of the Moon and i ts orbit around Earth, such that we
can only see one face of our natural satellite. In atomic phys ics, the spin-orbit coupling of electrons
orbiting around the nucleus gives rise to the atom’s fine stru cture (small shifts in its energy levels).
In condensed matter physics, spin-orbit effects are respons ible for exotic electronic phenomena in
semiconductors (topological insulators) and in supercond uctors without inversion symmetry. Al-
though spin-orbit coupling is ubiquitous in nature, it was n ot possible to control it in any area of
physics, until it was demonstrated in a breakthrough experi ment [1] that the spin of an atom could
be coupled to its center-of-mass motion by dressing twoatom ic spin states with apair of laser beams.
This unprecedented engineered spin-orbit coupling was pro duced in ultra-cold bosonic atoms, but
can also be created for ultra-cold fermionic atoms [1–3]. In anticipation of experiments, we develop
a theory for interacting fermions in the presence of spin-or bit coupling and Zeeman fields, and show
that many new superfluids phases, which are topological in na ture, emerge. Depending on values
of spin-orbit coupling, Zeeman fields, and interactions, in itially gapped s-wave superfluids acquire
p-wave,d-wave,f-wave and higher angular momentum components, which produc e zeros in the
excitation spectrum, rendering the superfluid gapless. Sev eral multi-critical points, which separate
topological superfluid phases from normal or non-uniform, a re accessible depending on spin-orbit
coupling, Zeeman fields or interactions, setting the stage f or the study of tunable topological super-
fluids.
PACS numbers: 03.75.Ss, 67.85.Lm, 67.85.-d
The effects of spin-orbit coupling in few body systems
like the Earth-Moon complex in astrophysics or the elec-
tron spin and its orbital motion around the nucleus in
isolated atoms of atomic physics are reasonably well un-
derstood due to the simplificity of these systems. How-
ever, in the setting of many identical particles, spin-orbit
effects have revealed quite interesting surprises recently
running from topological insulators in semiconductors [4]
to exotic superconductivity [5] and non-equillibrium ef-
fects [6] depending on the precise form of the spin-orbit
coupling. In atomic physics the coupling arises from the
interaction of the magnetic moment of the electron and
a magnetic field, present in the frame of electron, due
to the electric field of the nucleus. Similarly in con-
densed matter physics, the coupling arises from the mag-
netic moment mof electrons, which move in the back-
ground of ions. In the electron’s reference frame, these
ions are responsible for a magnetic field B, which de-
pends on the electron’s momentum kand couple to elec-
tron’s spin. The resulting spin-orbit coupling has the
formHSO=−m·B=−/summationtext
jhj(k)σj,whereσjrep-
resents the Pauli matrices and hj(k) describes the j-th
component ( j=x,y,z) of the effective magnetic field
vectorh. For some materials hcan take the Dressel-
haus [7] form hD(k) =vD(kyˆx+kxˆy),the Rashba [8]
formhR(k) =vR(−kyˆx+kxˆy),or more generally a lin-
ear combination of the two h⊥(k) =hD(k)+hR(k).In
all these situations the type of spin-orbit coupling cannot be changed arbitrarily and the magnitude can not
be tuned from weak to strong, making the experimental
control of spin-orbit effects very difficult.
Recently, however, it has been demonstrated experi-
mentally that spin-orbit coupling can be engineered in
a ultra-cold gas of bosonic atoms in their Bose-Einstein
condensatephase[1], whenapairofRamanlaserscreates
a coupling between two internal spin states of the atoms
and its center-of-mass motion (momentum). Thus far,
the type of spin-orbit field that has been created in the
laboratory [1] has the equal-Rashba-Dresselhaus (ERD)
formh⊥(k) =hERD(k) =vkxˆy, wherevR=vD=v/2.
Other forms of spin-orbit fields require additional lasers
and create further experimental difficulties [9]. In ultra-
coldbosonsthemomentum-dependent ERDcouplinghas
been created in conjunction with uniform Zeeman terms,
which are independent of momentum, along the z axis
(controlled by the Raman coupling Ω R), and along the
y-axis (controled by the detuning δ). The simultaneous
presence of hz,hyandhERD(k) leads to the Zeeman-
spin-orbit (ZSO) Hamiltonian
HZSO(k) =−hzσz−hyσy−hERD(k)σy
for an atom with center-of-mass momentum kand spin
basis| ↑/an}b∇acket∇i}ht,| ↓/an}b∇acket∇i}ht. The fields hz=−ΩR/2,hy=−δ/2 and
hERD=vkxˆycan be controlled independently, and thus
can be used as tunable parameters to explore the avail-
able phase space and to investigate phase transitions, as2
achieved in the experiment involving a bosonic isotope of
Rubidium (87Rb). Although current experiments have
focused on Bose atoms, there is no fundamental reason
that impeeds the realization of a similar set up for Fermi
atoms [1–3] designed to study fermionic superfluidity [3].
Considering possible experiments with fermionic atoms
such as6Li,40K, we discuss in this letter phase diagrams,
topological phase transitions, spectroscopic and thermo-
dynamic properties at zero and finite temperatures dur-
ing the evolution from BCS to BEC superfluidity in the
presence of controllable Zeeman and spin-orbit fields in
three dimensions.
To investigate artificial spin-orbit and Zeeman fields in
ultra-cold Fermi superfluids, we start from the Hamilto-
nian density
H(r) =H0(r)+HI(r), (1)
where the single-particle term is simply
H0(r) =/summationdisplay
αβψ†
α(r)
ˆKαδαβ−/summationdisplay
jˆhj(r)σj,αβ
ψβ(r).
(2)
Here,ˆKα=−∇2/(2m)−µαis the kinetic energyin refer-
ence to the chemical potential µα,ˆhj(r) is the combined
effective field including Zeeman and spin-orbit compo-
nents along the j-direction ( j=x,y,z), andψ†
α(r) are
creation operators for fermions with spin αat position
r. Notice that we allow the chemical potential µ↑to be
different from µ↓, such that the number of fermions N↑
with spin ↑may be different from the number of fermions
with spin ↓. The interaction term is
HI(r) =−gψ†
↑(r)ψ†
↓(r)ψ↓(r)ψ↑(r), (3)
wheregrepresents a contact interaction that can be
expressed in terms of the scattering length via the
Lippman-Schwinger relation V/g=−Vm/(4πas) +/summationtext
k1/(2ǫk).The introduction of the averagepairing field
∆(r)≡g/an}b∇acketle{tψ↓(r)ψ↑(r)/an}b∇acket∇i}ht ≈∆0and its spatio-temporal fluc-
tuationη(r,τ) produce a complete theory for superfluid-
ity in this system.
From now on, we focus on the experimental case where
a) the Raman detuning is zero ( δ= 0) indicating that
there is no component of the Zeeman field along the ydi-
rection; b) the Raman coupling Ω Ris non-zero meaning
that a Zeeman component along the zdirection exists,
that is,hz=−ΩR/2; and c) the spin-orbit field has com-
ponentshy(k) andhx(k) alongtheyandxdirections. To
start our discussion, we neglect fluctuations, and trans-
formH0(r) into momentum space as H0(k). Using the
basisψ†
↑(k)|0/an}b∇acket∇i}ht ≡ |k↑/an}b∇acket∇i}ht, ψ†
↓(k)|0/an}b∇acket∇i}ht ≡ |k↓/an}b∇acket∇i}ht,where|0/an}b∇acket∇i}htis
the vacuum state, the Fourier-transformed Hamiltonian
H0(k) becomes the matrix
H0(k) =K+(k)1+K−σz−hzσz−hy(k)σy−hx(k)σx,
Such matrix can be diagonalized in the helicity basis
Φ†
⇑(k)|0/an}b∇acket∇i}ht ≡ |k⇑/an}b∇acket∇i}ht,Φ†
⇓(k)|0/an}b∇acket∇i}ht ≡ |k⇓/an}b∇acket∇i}ht,where the spins⇑and⇓are aligned or antialigned with respect to the
effective magnetic field heff(k) =h/bardbl(k) +h⊥(k).Here,
K+(k) = (K↑+K↓)/2 =ǫk−µ+,is a measure of the
average kinetic energy ǫk=k2/2min relation to the
average chemical potential µ+= (µ↑+µ↓)/2.While
h⊥(k) =hx(k)ˆx+hy(k)ˆyis the spin-orbit field and
h/bardbl(k) = (hz−K−)ˆzis the effective Zeeman field, with
K−= (K↑−K↓)/2 =−µ−whereµ−= (µ↑−µ↓)/2 is
the internal Zeeman field due to initial population im-
balance, and hzis the external Zeeman field. When
there is no population imbalance the internal Zeeman
field isµ−= 0, and we have only hz. In general,
the eigenvalues of the Hamiltonian matrix H0(k) are
ξ⇑(k) =K+(k)−|heff(k)|andξ⇓(k) =K+(k)+|heff(k)|,
where|heff(k)|=/radicalbig
(µ−+hz)2+|h⊥(k)|2is the magni-
tude of the effective magnetic field, with the transverse
component being expressedin termsof the complex func-
tionh⊥(k) =hx(k) +ihy(k).In the limit where the in-
ternalµ−and external hzZeeman fields vanish and the
spin-orbit field is null ( h⊥= 0), the energies of the helic-
ity bands are identical ξ⇑(k) =ξ⇓(k) producing no effect
in the original energy dispersions [10].
When interactions are added to the problem, pair-
ing can occur within the same helicity band (intra-
helicity pairing) or between two different helicity bands
(inter-helicity pairing). This leads to a tensor order
parameter for superfluidity that has four components
∆⇑⇑(k) =−∆T(k)e−iϕ,corresponding to the helicity
projectionλ= +1; ∆ ⇑⇓(k) =−∆S(k),and ∆ ⇓⇑(k) =
∆S(k),corresponding to helicity projection λ= 0; and
∆⇓⇓(k) =−∆T(k)eiϕ,corresponding to helicity pro-
jectionλ=−1. The phase ϕ(k) is defined from
the amplitude-phase representation of the complex spin-
orbit fieldh⊥(k) =|h⊥(k)|eiϕ(k),while the amplitude
∆T(k) = ∆0|h⊥(k)|/|heff(k)|for helicities λ=±1aredi-
rectly proportional to the scalar order parameter ∆ 0and
to the relative magnitude of the spin-orbit field |h⊥(k)|
with respect to the magnitude of the effective magnetic
field|heff(k)|. Additionally, ∆ Thas the simple physi-
cal interpretation of being the triplet component of the
order parameter in the helicity basis, which is induced
by the presence of non-zero spin-orbit field h⊥, but van-
ishes when h⊥= 0. Analogously the amplitude ∆ S(k) =
∆0h/bardbl(k)/|heff(k)|for helicity λ= 0 are directly propor-
tional to the scalar order parameter ∆ 0and to the rela-
tive magnitude of the total Zeeman field h/bardbl(k) =µ−+hz
with respect to the magnitude of the effective magnetic
field|heff(k)|. Additionally, ∆ Shas the simple physical
interpretationofbeingthesingletcomponentofthe order
parameter in the helicity basis. It is interesting to note
the relation |∆T(k)|2+|∆S(k)|2=|∆0|2,which, for fixed
|∆0|, shows that as |∆S(k)|increases, |∆T(k)|decreases
and vice-versa. Such relation indicates that the singlet
and triplet channels are not separable in the presence of
spin-orbitcoupling. Furthermore, the orderparameterin
the triplet sector ∆ ⇑⇑(k) and ∆ ⇓⇓(k) contains not only
p-wave,but also f-waveand evenhigherodd angularmo-
mentum contributions, as long as the total Zeeman field3
µ−+hzis non-zero. Similarly, the orderparameterin the
singlet sector ∆ ⇑⇓(k) and ∆ ⇓⇑(k) contains not only only
s-wave, but also d-wave and even higher even angular
momentum contributions, as long as the total Zeeman
fieldµ−+hzis non-zero. Higher angular momentum
pairing in the helicity basis, occurs because the original
local (zero-ranged) interaction in the original ( ↑,↓) spin
basis is transformed into a finite-ranged interaction in
the helicity basis ( ⇑,⇓). In the limiting case of zero total
Zeemanfield µ−+hz= 0, the singletcomponentvanishes
(∆S(k) = 0), while the triplet component becomes inde-
pendent of momentum (∆ T(k) = ∆ 0), leading to order
parameter ∆ ⇑⇑(k) =−h∗
⊥(k), and ∆ ⇓⇓(k) =−h⊥(k)
which contains only p-wave contributions [11], since the
components of h⊥(k) depend linearly on momentum k.
The eigenvalues Ej(k) of the Hamiltonian including
the order parameter contribution emerge from the diago-
nalization ofa 4 ×4 matrix (see supplementary material).
The two eigenvalues for quasiparticles are
E1(k) =/radicalbigg/parenleftig
ξh−−/radicalig
ξ2
h++|∆S(k)|2/parenrightig2
+|∆T(k)|2,(4)
corresponding to the highest-energy quasiparticle band,
and
E2(k) =/radicalbigg/parenleftig
ξh−+/radicalig
ξ2
h++|∆S(k)|2/parenrightig2
+|∆T(k)|2,(5)
corresponding to the lowest-energy quasiparticle band,
while the eigenvalues for quasiholes are E3(k) =−E2(k)
for highest-energy quasihole band and E4(k) =−E1(k)
for the lowest-energy quasihole band. The energy ξh−=
[ξ⇑(k)−ξ⇓(k)]/2 is momentum-dependent, corresponds
to the average energy difference between the helicity
bands and can be written as ξh−=−|heff(k)|,while
the energy ξh+= [ξ⇑(k)+ξ⇓(k)]/2 is also momentum
dependent, corresponds to the averaged energy sum of
the helicity bands and can be written as ξh+=K+(k) =
ǫk−µ+.
There are a few important points to notice about the
excitation spectrum of this system. First, notice that
E1(k)> E2(k)≥0. Second, that the eigenergies are
symmetric about zero, such that we can regard quasi-
holes (negative energy solutions) as anti-quasiparticles.
Third, that only E2(k) can have zeros (nodal regions)
correspondingto the locus in momentum space satisfying
the following conditions: a) ξh−=−/radicalig
ξ2
h++|∆S(k)|2,
which corresponds physically to the equality between the
effective magnetic field energy |heff(k)|and the excita-
tion energy for the singlet component/radicalig
ξ2
h++|∆S(k)|2;
and b)|∆T(k)|= 0,corresponding to zeros of the triplet
component of the order parameter in momentum space.
SinceE2(k)< E1(k), and only E2(k) can have ze-
ros, the low energy physics is dominated by this ein-
genvalue. In the case of equal Rashba-Dresselhaus
(ERD) where h⊥(k) =v|kx|, zeros ofE2(k) can oc-
cur whenkx= 0, leading to the following cases: (a)two possible lines (rings) of nodes at ( k2
y+k2
z)/(2m) =
µ++/radicalbig
(µ−+hz)2−|∆0|2for the outer ring, and ( k2
y+
k2
z)/(2m) =µ+−/radicalbig
(µ−+hz)2−|∆0|2forthe innerring,
when (µ−+hz)2− |∆0|2>0; (b) doubly-degenerate
line of nodes ( k2
y+k2
z)/(2m) =µ+forµ+>0, doubly-
degenerate point nodes for µ+= 0, or no-line of nodes
forµ+<0, when (µ−+hz)2− |∆0|2= 0; (c) no
line of nodes when ( µ−+hz)2− |∆0|2<0. In ad-
dition, case (a) can be refined into cases (a2), (a1)
and (a0). In case (a2), two rings indeed exist pro-
vided that µ+>/radicalbig
(µ−+hz)2−|∆0|2. However, the
inner ring disappears when µ+=/radicalbig
(µ−+hz)2−|∆0|2.
In case (a1), there is only one ring when |µ+|</radicalbig
(µ−+hz)2−|∆0|2,In case (a0), the outer ring dis-
appears at µ+=−/radicalbig
(µ−+hz)2−|∆0|2, and forµ+<
−/radicalbig
(µ−+hz)2−|∆0|2no rings exist.
We choose our momentum, energy and velocity scales
through the Fermi momentum kF+defined from the to-
tal density of fermions n+=n↑+n↓=k3
F+/(3π2).This
choice leads to the Fermi energy ǫF+=k2
F+/2mand to
the Fermi velocity vF+=kF+/m, as energy and veloc-
ity scales respectively. In Fig. 1, we show the phase
diagram of Zeeman field hz/ǫF+versus chemical poten-
tialµ+/ǫF+describing possible superfluid phases accord-
ing to their quasiparticle excitation spectrum. We la-
bel the uniform superfluid phases with zero, one or two
rings of nodes as US-0, US-1, and US-2, respectively.
Non-uniform (NU) phases also emerge in regions where
uniform phases are thermodynamically unstable. The
US-2/US-1 phase boundary is determined by the condi-
tionµ+=/radicalbig
(µ−+hz)2−|∆0|2, when|µ−+hz|>|∆0|;
the US-0/US-2 boundary is determined by the Clogston-
like condition |(µ−+hz)|=|∆0|whenµ+>0, where
the gapped US-0 phase disappears leading to the gap-
less US-2 phase; and the US-0/US-1 phase boundary
is determined by µ+=−/radicalbig
(µ−+hz)2−|∆0|2, when
|µ−+hz|>|∆0|. Furthermore, with the US-0 bound-
aries, a crossover line between an indirectly gapped and
a directly gapped US-0 phase occurs at µ+= 0. Lastly,
some important multi-critical points arise at the inter-
sections of phase boundaries. First the point µ+= 0 and
|(µ−+hz)|=|∆0|corresponds to a tri-critical point for
phasesUS-0, US-1, and US-2. Second, the point |∆0|= 0
andµ+=|(µ−+hz)|corresponds to a tri-critical point
for phases N, US-1 and US-2. In the limit where both
µ−andhzvanish no phase transitions take place and the
problem is reduced to a crossover [12–14].
In the US-1 and US-2 phases near the zeros of E2(k),
quasiparticles have linear dispersion and behave as Dirac
fermions. Such change in nodal structures is associated
with bulk topological phase transitions of the Lifshitz
classasnotedfor p-wave[15]and d-wave[16,17]superflu-
ids. Such Lifshitz topological phase transitions are possi-
ble here because the spin-orbit coupling field induces the
triplet component of the order parameter ∆ T(k). The
loss of nodal regions correspond to annihilation of Dirac4
a
N
NUS-1
US-1 NU
NU hz / !F+
"+ / !F+ -2 -1 0 10123
-3 -2 -1 b
-2 -1 0 10123
-3 -2 -1 N
NUS-1
US-1 Indirect
US-0 Direct
US-0
"+ / !F+ Indirect
US-0 NU
NU hz / !F+
Direct
US-0 US-2
US-2
FIG. 1: Phase diagram of Zeeman field hz/ǫF+versus chem-
ical potential µ+/ǫF+for a)v/vF+= 0 and b) v/vF+= 0.28
identifying uniform superfluid phases US-0 (gapped), US-1
(gapless with one ring of nodes), and US-2 (gapless with
two-rings of nodes). The NU region corresponds to unsta-
ble uniform superfluids which may include phase separation
and/or a modulated superfluid (supersolid). Solid lines rep -
resent phase boundaries, while the dashed line represents t he
crossover from the direct-gap to the indirect-gap US-0 phas e.
quasiparticles with opposite momenta, which lead to the
disappearance of rings. The transition from phase US-
2 to indirect gapped US-0 occurs through the merger of
the two-rings at the phase boundary followed by the im-
mediate opening of the indirect gap at finite momentum.
However, the transition from phase US-2 to US-1 corre-
sponds to the disappearance of the inner ring through
the origin of momenta, similarly the transition from US-
1 to the directly gappped US-0 corresponds to the dis-
appearance of the last ring also through the origin of
momenta. In the case of Rashba-only coupling rings
of nodes are absent and it is possible to have at most
nodal points [18, 19]. The last two phase transitions
are special because the zero-momentum quasiparticles at
thesephaseboundariescorrespondtotrueMajoranazero
energy modes if the phase ϕ(k) of the spin-orbit field
h⊥(k) =|h⊥(k)|eiϕ(k)and the phase θ(k) of the order
parameter ∆ 0=|∆0|eiθ(k)have opposite phases at zero
momentum: ϕ(0) =−θ(0) [mod(2π)]. This can be seen
from an analysis of the quasiparticle eigenfunction
Φ2(k) =u1(k)ψk↑+u2(k)ψk↓+u3(k)ψ†
−k↑+u4(k)ψ†
−k↓
corresponding to the eigenvalue E2(k). The emergence
ofzero-energyMajoranafermionsrequiresthequasiparti-
cle to be its own anti-quasiparticle: Φ†
2(k) = Φ2(k). This
canonlyhappenatzeromomentum k=0,wheretheam-
plitudesu1(0) =u∗
3(0) andu2(0) =u∗
4(0). Such require-
ment leads to the conditions µ2
+= (µ−+hz)2+|∆0|2,
andϕ(0) =−θ(0) [mod(2π)], showing that Majorana
fermions can exist only at the US-0/US-1 and US-2/US-
1 phase boundaries. It is important to emphasize that
the Majorana fermions found here exist in the bulk, and
thus their emergence or disappeareance affect bulk ther-modynamic properties, unlike Majorana fermions found
at the edge (surfaces) of topological insulators and some
topological superfluids. The common ground between
bulk and surface Majorana fermions is that both exist at
boundaries: the bulk Majorana zero-energy modes may
exist at the phase boundaries between two topologically
distinct superfluid phases, while surface Majorana zero-
energy modes may exist at the spatial boundaries of a
topologically non-trivial superfluid.
It is evident that the transition between different su-
perfluid phases occurs without a change in symmetry in
the orderparameter∆ 0, and thus violatesthe symmetry-
based Landau classification of phase transitions. In the
presentcase, the simultaneousexistenceofspin-orbitand
Zeeman fields (internal or external) couple the singlet
∆S(k) and triplet ∆ T(k) channels and all the super-
fluid phases US-0, US-1 and US-2 just have different
weights from each order parameter component. How-
ever a finer classification based on topological charges
can be made via the construction of topological invari-
ants. Since the superfluid phases US-0, US-1, US-2 are
characterized by different excitation spectra correspond-
ing to the eigenvalues of the Hamiltonian matrix includ-
ing interactions H(k), we can use the resolvent matrix
R(ω,k) = [−ω1+H(k)]−1and the methods of algebraic
topology [20] to construct the topological invariant
ℓ=/integraldisplay
DdSγ
24π2ǫµνλγTr/bracketleftbig
ΛkµΛkνΛkλ/bracketrightbig
,
whereΛkµ=R∂kµR−1.The topological invariant is
ℓ= 0 in the gapped US-0 phase, is ℓ= 1 in the gap-
less US-1 phase and ℓ= 2 in the gapless US-2 phase,
showing that, for ERD spin-orbit coupling, ℓcounts the
number of rings of zero-energy excitations in each super-
fluid phase. The integral above has a hyper-surface mea-
suredSγandadomain Dthat enclosesthe regionofzeros
ofω=Ej(k) = 0. Here µ,ν,λ,γ run from 0 to 3, and kµ
has components k0=ω,k1=kx,k2=ky, andk3=kz.
The topological invariant measures the flux of the four-
dimensional vector Fγ=ǫµνλγTr/bracketleftbig
ΛkµΛkνΛkλ/bracketrightbig
/24π2,
through a hypercube including the singular region of the
resolvent matrix R(ω,k), much in the same way that the
flux of the electric field Ein Gauss’ law of classical elec-
tromagnetism measures the electric charge qenclosed by
a Gaussian surface:/contintegraltext
dS·(ǫ0E) =q. Thus, the topologi-
cal invariant defined above defines the topological charge
of fermionic excitations, in the same sense as Gauss’ law
for the electric flux defines the electric charge.
A full phase diagram can be constructed only upon
verification of thermodynamic stability of all the pro-
posed phases. For this purpose it becomes imper-
ative to investigate the maximum entropy condition
(see supplementary material). Independent of any
microscopic approximations, the necessary and suffi-
cient conditions for thermodynamic stability of a given
phase are: positive isovolumetric heat capacity CV=
T(∂S/∂T)V,{Nα}≥0; positive chemical susceptibility
matrixξαβ= (∂µα/∂Nβ)T,V,i.e, eigenvalues of the5
matrix [ξ] are both positive; and positive bulk mod-
ulusB= 1/κTor isothermal compressibility κT=
−V−1(∂V/∂P)T,{Nα}.Using these conditions, we con-
struct the full phase diagramdescribed in Fig. 1 for equal
Rasha-Dresselhaus (ERD) spin-orbit coupling. The re-
gions, where the uniform superfluid phases are unsta-
ble are labeled by the abbreviation NU to indicate that
non-uniform phases such as phase separation or modu-
lated superfluid (supersolid) may emerge. In Fig. 2, we
showthephasediagramofZeemanfield hz/ǫF+versusin-
teraction parameter 1 /(kF+as), for population balanced
fermions, where the number of spin-up fermions N↑is
equal to the number of spin-down fermions N↓.
v / v_F = 0.141
b
N
0 1 2 3 4-1 -2 2468
0hz / !F+
US-2 Indirect US-0 Direct
US-0 US-1
1/k F+ asNU
v / v_F = 0.283
c
N
US-1 Direct
US-0
US-2 Indirect US-0
0 1 2 3 4-1 -2 2468
0hz / !F+
1/k F+ asNU v / v_F = 0.424
0 1 2 3 4-1 -2 2468
0d
NDirect
US-0
US-2 Indirect US-0 hz / !F+
1/k F+ asUS-1
NU v / v_F = 0
a
N
0 1 2 3 4-1 -2 2468
0hz / !F+
Indirect US-0 Direct
US-0
1/k F+ asNU US-1
FIG.2: Phasediagram ofZeemanfield hz/ǫF+versusinterac-
tion 1/(kF+as) showing uniform superfluid phases US-0, US-
1, and US-2, and non-uniform (NU) region for a) v/vF+= 0;
b)v/vF+= 0.14; c)v/vF+= 0.28; d)v/vF+= 0.56.
Solid lines are phase boundaries, the dashed line indicates
a crossover from the indirect- to direct-gapped US-0.
Since these superfluid phases exhibit major changes in
momentum-frequency space as evidenced by their single
particle excitation spectrum, it is important to explore
additional spectroscopic quantitities to characterize fur-
ther the nature of these phases and the phase transitions
between them. An important quantity is the 4 ×4 resol-
vent matrix
R(iω,k) =/parenleftbiggG(iω,k)F(iω,k)
F†(iω,k)G(iω,k)/parenrightbigg
,(6)
from where the spectral density Aα(ω,k) =
−(1/π)ImGαα(iω=ω+iδ,k) for spin α=↑,↓can
be extracted. The spectral function Aα(ω,k) in the
plane of momenta ky-kzwithkx= 0 and frequency
ω= 0 reveals the existence of rings of zero-energy
excitations in the US-1 and US-2 phases. The density of
statesDα(ω) =/summationtext
kAα(ω,k) for spinαas a function of
frequencyωis also an important spectroscopic quantitywhich is shown in Fig. 3 along with excitation spectra
Ej(k) for phases US-1 and US-2 at fixed ERD spin-orbit
couplingv/vF+= 0.28. The parameters used for phase
US-1 arehz/ǫF+= 0.5 and 1/(kF+as) =−0.4, while for
phase US-2 they are hz/ǫF+= 2.0 and 1/(kF+as) = 1.0.
Notice that, even though the excitation spectrum Ej(k)
is symmetric, the coherence factors appearing in the
matrixGare not, such that the density of states Dα(ω)
is not an even function of ω, and thus it is not particle-
hole symmetric. The main feature of Dα(ω) at low
frequencies is the linear behavior due to the existence
of Dirac quasiparticles and quasiholes in the US-1 and
US-2 phases, which are absent in the direct-gap and the
indirect-gap US-0 phases. The peaks and structures in
Dα(ω) mostly emerge due to the maxima and minima of
Ej(k). Notice that for finite Zeeman field hz, the density
of states D↑(ω)/ne}ationslash=D↓(ω) because the induced population
imbalanceP= (N↑−N↓)/(N↑+N↓) is non-zero. For
the US-2 case shown in Fig. 3b, the induced population
imbalanceP≪1 sincehz/ǫF+is small, while for the
US-1 case shown in Fig. 3e, P≈1 as the spins are
almost fully polarized since hz/ǫF+is large.
!!"# $$"# %!!"& !!"' !!"' !"&
!!"# $$"# %!!"& !!"' !!"' !"&
!"#$%!!&' !!&% !!&% !&'
!!"# $$"# %!!"& !!"' !!"' !"&
!!"# $$"# %!!"& !!"' !!"' !"&
!!"# $$"# %!!"& !!"' !!"' !"& |k x| / k F+ Ej / !F+
|k y| / k F+ Ej / !F+
|k x| / k F+ Ej / !F+
|k y| / k F+ Ej / !F+ D( ") !F+ " / !F+
D( ") !F+ " / !F+ a b c
f e d
FIG. 3: Energy spectrum and density of states in phase US-2
areshown ina), b), c)for hz/ǫF+= 0.5and1/(kF+as) =−0.4
and in phase US-1 are shown in d), e), f) for hz/ǫF+= 2.0
and 1/(kF+as) = 1.0. Energies Ej(kx,0,0) versus |kx|in a)
and d); frequency ωversus density of states D↑(ω) (dashed),
D↓(ω) (dot-dashed), and their sum D(ω) (solid) in b) and e);
energies Ej(0,ky,0) versus |ky|in c) and f).
In summary, we have discussed the effects of spin-orbit
and Zeeman fields in ultra-cold Fermi superfluids, ob-
tained the phase diagrams of Zeeman field versus inter-
action parameter or versus chemical potential, and iden-
tified several bulk topological phase transitions between
gapped and gapless superfluids as well as a variety of
multi-critical points. We haveshown that the presenceof
simultaneousZeeman and spin-orbitfields induces higher6
angular momentum pairing, as manifested in the emer-
gence of momentum dependence ofthe singlet and triplet
components of order parameter expressed in the helicity
basis. Finally, we have characterized topological phases
and phase transitions between them through their exci-
tation spectra (existence of Dirac quasiparticles or Majo-
rana zero-energy modes), topological charges, and spec-
troscopic and thermodynamic properties, such as densityof states and isothermal compressibility.
Acknowledgments
We thank ARO (W911NF-09-1-0220) for support.
[1] Y. J. Lin, K. Jimenez-Garcia, and I. B. Spielman, Nature
471, 83-86 (2011).
[2] X. J. Liu, M. F. Borunda, X. Liu, and J. Sinova, Phys.
Rev. Lett. 102, 046402 (2009).
[3] M. Chapman and C. S´ a de Melo, Nature 471, 41-42
(2011).
[4] C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 146802
(2005).
[5] L. P. Gor’kov and E. I. Rashba, Phys. Rev. Lett. 87,
037004 (2001).
[6] T. D. Stanescu, C. Zhang, and V. Galitski, Phys. Rev.
Lett.99110403 (2007).
[7] G. Dresselhaus, Phys. Rev. 100, 580 (1955).
[8] Y. A. Bychkov and E. I. Rashba, J. Phys. C 17, 6039
(1984).
[9] J. Dalibard, F. Gerbier, G. Juzeliunas, and P. Ohberg,
arXiv:1008.5378 (2010).
[10] J. P. Vyasanakere, S. Zhang, and V. B. Shenoy,
arXiv:1104.5633 (2011).
[11] C. Zhang, S. Tewari, R. M. Lutchyn, and S. Das Sarma,
Phys. Rev. Lett. 101160401 (2008).
[12] Z.-Q. Yu and H. Zhai, arXiv:1105.2250 (2011).
[13] H. Hu, L. Jiang, X. Jiu, and H. Pu, arXiv:1105.2488
(2011).
[14] L. Han and C. A. R. S´ a de Melo, arXiv:1106.3613 (2011).
[15] G. E. Volovik, World Scientific, Singapore (1992).
[16] R. D. Duncan and C. A. R. S´ a de Melo, Phys. Rev. B
62, 9675 (2000).
[17] S. S. Botelho and C. A. R. S´ a de Melo, Phys. Rev. B 71,
134507 (2005).
[18] M. Gong, S. Tewari, and C. Zhang, arXiv:1105.1796
(2011).
[19] M. Iskin and A. L. Subasi, arXiv:1106.0473 (2011).
[20] M. Nakahara, Adam Hilger, Bristol (1990).
I. ARTIFICIAL SPIN-ORBIT COUPLING IN
ULTRA-COLD FERMI SUPERFLUIDS:
(SUPPLEMENTARY MATERIAL)
The method used to study the spin-orbit and Zeeman
effects in ultra-cold Fermi superfluids is the functional
integral method and its saddle-point approximation in
conjunction with fluctuation effects. To describe the
thermodynamic phases and the corresponding phase dia-
gram in terms of the interactions, Zeeman and spin-orbit
fields, we calculate partition function at temperature TZ=/integraltext
D[ψ,ψ†]exp/parenleftbig
−S[ψ,ψ†]/parenrightbig
with action
S[ψ,ψ†] =/integraldisplay
dτdr/bracketleftigg/summationdisplay
αψ†
α(r,τ)∂
∂τψα(r,τ)+H(r,τ)/bracketrightigg
,
where the Hamiltonian density is given in Eq. (1).
Using the standard Hubbard-Stratanovich transfor-
mation that introduces the pairing field ∆( r,τ) =
g/an}b∇acketle{tψ↓(r,τ)ψ↑(r,τ)/an}b∇acket∇i}htand integrating over the fermion vari-
ables lead to the effective action
Seff=/integraldisplay
dτdr/bracketleftbigg|∆(r,τ)|2
g−T
2VlndetM
T+/tildewideK+δ(r−r′)/bracketrightbigg
,
where/tildewideK+= (/tildewideK↑+/tildewideK↓)/2.The matrix Mis
M=
∂τ+/tildewideK↑−h⊥0−∆
−h∗
⊥∂τ+/tildewideK↓∆ 0
0 ∆†∂τ−/tildewideK↑h∗
⊥
−∆†0h⊥∂τ−/tildewideK↓
,(7)
whereh⊥=hx−ihycorresponds to the transverse com-
ponent of the spin-orbit field, hzto the parallel com-
ponent with respect to the quantization axis z,/tildewideK↑=
ˆK↑−hz, and/tildewideK↓=ˆK↓+hz.
To make progress, we use the saddle point approxi-
mation ∆( r,τ) = ∆ 0+η(r,τ),and write M=Msp+
Mf. The matrix Mspis obtained via the saddle point
∆(r,τ)→∆0which takes M→Msp, and the fluctua-
tion matrix Mf=M−Mspdepends only on η(r,τ) and
its Hermitian conjugate. Thus, we write the effective ac-
tion asSeff=Ssp+Sf. The first term is
Ssp=V
T|∆0|2
g−1
2/summationdisplay
k,iωn,jln/bracketleftbigg−iωn+Ej(k)
T/bracketrightbigg
+/summationdisplay
k/tildewideK+
T,
in momentum-frequency coordinates ( k,iωn), where
ωn= (2n+1)πT. Here,Ej(k) are the eigenvalues of
Hsp=
/tildewideK↑(k)−h⊥(k) 0 −∆0
−h∗
⊥(k)/tildewideK↓(k) ∆ 0 0
0 Ơ
0−/tildewideK↑(−k)h∗
⊥(−k)
−∆†
00h⊥(−k)−/tildewideK↓(−k)
,
(8)
which describes the Hamiltonian of elemen-
tary excitations in the four-dimensional basis7
Ψ†=/braceleftig
ψ†
↑(k),ψ†
↓(k),ψ↑(−k),ψ↓(−k)/bracerightig
.The fluctu-
ation action is
Sf=/integraldisplay
dτdr/bracketleftbigg|η(r,τ)|2
g−T
2Vlndet/parenleftbig
1+M−1
spMf/parenrightbig/bracketrightbigg
.
The spin-orbit field is h⊥(k) =hR(k) +hD(k),
wherehR(k) =vR(−kyˆx+kxˆy) is of Rashba-type and
hD(k) =vD(kyˆx+kxˆy) is of Dresselhaus-type, has
magnitude |h⊥(k)|=/radicalig
(vD−vR)2k2y+(vD+vR)2k2x.
For Rashba-only (RO) ( vD= 0) and for equal Rashba-
Dresselhaus(ERD) couplings( vR=vD=v/2), the mag-
nitude of the transverse fields are |h⊥(k)|=vR/radicalig
k2x+k2y
(vR>0) andh⊥(k) =v|kx|(v>0), respectively.
The Hamiltonian in the helicity basis Φ = UΨ, where
Uis the unitary matrix that diagonalizes the Hamilto-
nian in the normal state, is
/tildewideHsp(k) =
ξ⇑(k) 0 ∆ ⇑⇑(k) ∆⇑⇓(k)
0ξ⇓(k) ∆⇓⇑(k) ∆⇓⇓(k)
∆∗
⇑⇑(k) ∆∗
⇑⇓(k)−ξ⇑(k) 0
∆∗
⇑⇓(k) ∆∗
⇓⇓(k) 0 −ξ⇓(k)
.
The components of the order parameter in the helic-
ity basis are given by ∆ ⇑⇑(k) = ∆ T(k)e−iϕk,and
∆⇓⇓(k) =−∆T(k)eiϕkfor the triplet channel and by
∆⇑⇓(k) =−∆S(k) and ∆ ⇓⇑(k) = ∆ S(k) for the sin-
glet channel. The eigenvalues of Hsp(k) for quasiparti-
clesE1(k),E2(k) are listed in Eqs. (4) and (5), while
the eigenvalues for quasiholes are E3(k) =−E2(k), and
E4(k) =−E1(k).
The thermodynamic potential is Ω = Ω sp+Ωf, where
Ωsp=V|∆0|2
g−T
2/summationdisplay
k,jln{1+exp[−Ej(k)/T]}+/summationdisplay
k¯K+,
with¯K+=/bracketleftig
/tildewideK↑(−k)+/tildewideK↓(−k)/bracketrightig
/2 is the saddle
point contribution and Ω f=−TlnZf, withZf=/integraltext
D[¯η,η]exp[−Sf(¯η,η)] is the fluctuation contribution.
The order parameter is determined via the minimization
of Ωspwith respect to |∆0|2, leading to
V
g=−1
2/summationdisplay
k,jnF[Ej(k)]∂Ej(k)
∂|∆0|2, (9)
wherenF[Ej(k)] = 1/(exp[Ej(k)/T] + 1) is the Fermi
function for energy Ej(k). The contact interaction gis
expressed in terms of the scattering parameter asvia the
Lippman-Schwinger relation discussed in the main text.
The total number of particles N+=N↑+N↓is defined
fromthethermodynamicrelation N+=−(∂Ω/∂µ+)T,V,
and can be written as
N+=Nsp+Nf. (10)
The saddle point contribution is
Nsp=−/parenleftbigg∂Ωsp
∂µ+/parenrightbigg
T,V=1
2/summationdisplay
k
1−/summationdisplay
jnF[Ej(k)]∂Ej(k)
∂µ+
,and the fluctuation contribution is Nf=
−(∂Ωf/∂µ+,)T,Vleading to
Nf=T
Zf/integraldisplay
D[¯η,η]exp[−Sf(¯η,η)]/parenleftbigg
−∂Sf(¯η,η)
∂µ+/parenrightbigg
,
with the partial derivative being
∂SF(¯η,η)
∂µ+=−T
2VTr/bracketleftbigg/parenleftbig
1+M−1
spMf/parenrightbig−1∂
∂µ+/parenleftbig
M−1
spMf/parenrightbig/bracketrightbigg
.
Knowledge of the thermodynamic potential Ω, of the
order parameter Eq. (9) and number Eq. (10) provides
a complete theory for spectroscopic and thermodynamic
properties of attractive ultra-cold fermions in the pres-
ence of Zeeman and spin-orbit fields. Representative
Saddle point solutions for chemical potential µ+and or-
der parameter amplitude |∆0|as a function of 1 /(kF+as)
in the equal Rashba-Dresselhaus (ERD) case ( v/vF+=
0.28) are shown in Fig. 4 for hz/ǫF+= 0,0.5,1.0,2.0.
These parameters are used to obtain the phase diagrams
described in Figs. 1 and 2 in combination with an anal-
ysis of the excitation spectrum Ej(k) given in Eqs. (4)
and (5) and the thermodynamic stability conditions for
all the uniform superfluid phases: directly or indirectly
gapped superfluid with zero nodal rings (US-0); gapless
superfluid with one ring of nodes (US-1); and gapless
superfluid with two rings of nodes (US-2).
!!!"#"!$%#"!$
!!!"#"!$%!%!!#!
1/k F+ as 1/k F+ as!+ / "F+
|#0| / "F+ b a
FIG. 4: a) Chemical potential µ+/ǫF+and b) order pa-
rameter amplitude |∆0|/ǫF+versus interaction parameter
1/(kF+as) for spin-orbit parameter v/vF+= 0.28 and val-
ues of the Zeeman field hz/ǫF+= 0 (solid); hz/ǫF+= 0.5
(dashed); hz/ǫF+= 1.0 (dotted); and hz/ǫF+= 2.0 (dot-
dashed).
A thermodynamic stability analysis of all proposed
phases can be performed by investigating the maximum
entropy condition. The total change in entropy due to
thermodynamic fluctuations, irrespective to any approx-
imations imposed on the microscopic Hamiltonian, can
be written as
∆Stot=−1
2T(∆T∆S−∆P∆V+∆µα∆Nα),
where the repeated αindex indicates summation, and
the condition ∆ Stot≤0 guarantees that the entropy is
maximum. Considering the entropy Sto be a function
of temperature T, number of particles Nαand volume8
V, we can elliminate the fluctuations ∆ S, ∆P, and ∆µα
in favor of fluctuations ∆ T, ∆Vand ∆Nα, and show
that the fluctuations ∆ Tare statistically independent of
∆Nαand ∆V, while fluctuations ∆ Nαand ∆Vare not.
The first condition for thermodynamic stability leads
to the requirement that the isovolumetric heat capacity
CV=T(∂S/∂T)V,{Nα}≥0.Additional conditions are
directly related to number ∆ Nαand volume ∆ Vfluctu-
ations. They require the chemical susceptibility matrix
ξαβ= (∂µα/∂Nβ)T,Vto be positive definite, i.e, that
its eigenvalues are both positive. This is guaranteed by
det[ξ] =ξ↑↑ξ↓↓−ξ↑↓ξ↓↑>0 andξ↑↑>0. The last con-
dition for thermodynamic stability is that the bulk mod-
ulusB= 1/κTor the isothermal compressibility κT=
−V−1(∂V/∂P)T,{Nα},are positive. Since the number
∆Nαand volume ∆ Vfluctuations are not statistically
independent, the bulk modulus is related to the matrix
[ξ] viaV/κT=N2
↑ξ↑↑+N↑N↓ξ↑↓+N↓N↑ξ↓↑+N2
↓ξ↓↓.
The positivity of the volumetric specific heat CV, chemi-
cal susceptibility matrix [ ξ] and bulk modulus B= 1/κT
are the necessary and sufficient conditions for thermody-
namic stability, which must be satisfied irrespective of
approximations used at the microscopic level.
!!!"#"!$%#%&"! "'
!!!"#"!$%#%&"! "' !!!"#"!$%#%&"! "'
!!!"#"!$%#%&"! "' !T "F+
!T "F+ !T "F+
!T "F+ 1/k F+ as 1/k F+ as
1/k F+ as 1/k F+ asd cb a
0.5 11.5 025 50
!!"# $!#! !#! !!"# !!"$# !"% !#! !#!
FIG. 5: Isothermal compressibility ¯ κT= (N2
+)V−1κT=
(∂N+/∂µ+)T,Vin units of 3 N+/(4ǫF+) versus interaction
1/(kF+as) at spin-orbit coupling v/vF+= 0.28 for the val-
ues of the Zeeman field a) hz/ǫF+= 0; b) hz/ǫF+= 0.5;
c)hz/ǫF+= 1.0; and d) hz/ǫF+= 2.0. Insets show regions
where the compressibility is large.
Further characterization of phases US-0, US-1 and US-2 is made via thermodynamic prop-
erties such as the isothermal compressibility
κT= (V/N2
+)(∂N+/∂µ+)T,V,which is shown in
Fig. 5 versus 1 /(kF+as) for the values of the Zee-
man field hz/ǫF+= 0,0.5,1.0,2.0 and spin-orbit
couplingv/vF+= 0.28. Notice the negative re-
gions ofκTindicating that the uniform superfluid
phases are unstable, and its discontinuities at phase
boundaries. The normal state compressibility κTor
¯κT= (N2
+)V−1κT= (∂N+/∂µ+)T,Vcan be obtained
analytically for arbitrary Zeeman hzand spin-orbit
parametervin the BCS limit where 1 /kF+as→ −∞as
¯κT=3N+
4ǫF+/summationdisplay
j=±/bracketleftbigg
Aj+/bracketleftbigg
˜µ+−A2
j+/radicalig
˜h2z+2˜vA2
j/bracketrightbigg∂Aj
∂˜µ+/bracketrightbigg
,
(11)
where the auxiliary function Ajis
A±=/radicaligg
(˜µ++ ˜v)±/radicalbigg
(˜µ++ ˜v)2−/parenleftig
˜µ2
+−˜h2z/parenrightig
and its derivative is
∂A±
∂˜µ+=/bracketleftbigg
1±˜v//radicalig
(˜µ++ ˜v)2−(˜µ2
+−˜h2z)/bracketrightbigg
/(2A±)
with ˜µ+=µ+/ǫF+,˜hz=hz/ǫF+, and ˜v=v/(2ǫF+).
Notice that, as hz→0 and ˜v→0,A±→√˜µ+and
¯κT→(3N+)/(2ǫF+) is reduced to the standard result,
since ˜µ+→1. In addition, κTor ¯κTcan be obtained
analytically in the BEC limit where 1 /kF+as→+∞.
Whenhzandvare zero, then
¯κT=3N+
2ǫF+π
kF+as(12)
can also be written in terms of bosonic properties
1
V/parenleftbigg∂N+
∂µ+/parenrightbigg
T,V=1
π/parenleftbiggmB
aB/parenrightbigg
, (13)
wheremB= 2mis the boson mass and aB= 2asin
the boson-boson interaction. In the case where hz/ne}ationslash= 0
andv/ne}ationslash= 0, a similar expression can be derived for
V−1(∂N+/∂µ+)T,Vbut the effective boson mass mB=
2mf(hz,v), and the effective boson-boson interaction
aB= 2asg(hz,v) are now functions of hzandv. Notice
that the ratio mB/aBin the BEC limit can be directly
extracted from the behavior of ¯ κTfor large 1/(kF+as). |
2201.06265v2.Spin_orbit_coupled_superconductivity_with_spin_singlet_non_unitary_pairing.pdf | Spin-orbit-coupled superconductivity with spin-singlet non-unitary pairing
Meng Zeng,1Dong-Hui Xu,2, 3Zi-Ming Wang,2, 3and Lun-Hui Hu4, 5,
1Department of Physics, University of California, San Diego, California 92093, USA
2Department of Physics and Chongqing Key Laboratory for Strongly Coupled Physics,
Chongqing University, Chongqing 400044, China
3Center of Quantum Materials and Devices, Chongqing University, Chongqing 400044, China
4Department of Physics, the Pennsylvania State University, University Park, PA, 16802, USA
5Department of Physics and Astronomy, University of Tennessee, Knoxville, Tennessee 37996, USA
The gap functions for a single-band model for unconventional superconductivity are distinguished
by their unitary or non-unitary forms. Here we generalize this classication to a two-band super-
conductor with two nearly degenerate orbitals. We focus on spin-singlet pairings and investigate the
eects of the atomic spin-orbit coupling (SOC) on superconductivity which is a driving force behind
the discovery of a new spin-orbit-coupled non-unitary superconductor. Multi-orbital eects like or-
bital hybridization and strain induced anisotropy will also be considered. The spin-orbit-coupled
non-unitary superconductor has three main features. First, the atomic SOC locks the electron
spins to be out-of-plane, leading to a new Type II Ising superconductor with a large in-plane upper
critical eld beyond the conventional Pauli limit. Second, it provides a promising platform to re-
alize the topological chiral or helical Majorana edge state even without external magnetic elds or
Zeeman elds. More surprisingly, a spin-polarized superconducting state could be generated by spin-
singlet non-unitary pairings when time-reversal symmetry is spontaneously broken, which serves as
a smoking gun to detect this exotic state by measuring the spin-resolved density of states. Our
work indicates the essential roles of orbital-triplet pairings in both unconventional and topological
superconductivity.
I. INTRODUCTION
In condensed matter physics, research on unconven-
tional superconductivity [1, 2] remains a crucial topic
and continues to uncover new questions and challenges
in both theory and experiment, since the discovery of the
heavy-fermion superconductors (SCs) [3] and the d-wave
pairing states in high-temperature cuprate SCs [4{7]. In
addition to the anisotropic gap functions (e.g., p;d;f;g -
wave...), the sublattice or orbital-dependent pairings [8{
10] are shown to be an alternative avenue to real-
ize unconventional SCs. They might be realized in
multi-orbital correlated electronic systems, whose candi-
date materials include iron-based SCs [11{22], Cu-doped
Bi2Se3[23, 24], half-Heusler compounds [25{34], and pos-
sibly Sr 2RuO 4[35{40] etc. In particular, considering
the atomic orbital degrees of freedom, the classication
of unconventional pairing states could be signicantly
enriched. Among them, SCs with spontaneous time-
reversal symmetry (TRS) breaking is of special interest,
in which two mutually exclusive quantum phenomena,
spin magnetism, and superconductivity may coexist with
each other peacefully[41{46].
On the other hand, the orbital multiplicity could also
give rise to non-unitary pairings, which again include
both time-reversal breaking (TRB) and time-reversal
invariant (TRI) pairings. Very recently, prior studies
have demonstrated the existence of spin-singlet non-
unitary pairing states that break the inversion symme-
try in Dirac materials [9]. One aim of this work is the
hu.lunhui.zju@gmail.comgeneralization of unitary and non-unitary gap functions
in a two-band SC while preserving inversion symme-
try, which is possible exactly due to the multi-orbital
degrees of freedom [47]. We focus on a system with
two nearly degenerate orbitals and nd that the non-
unitary pairing state is generally a mixed superconduct-
ing state with both orbital-independent pairings and
orbital-dependent pairings. Recently, the interplay be-
tween orbital-independent pairings and spin-orbit cou-
pling (SOC) has been shown to demonstrate the intrigu-
ing phenomenon of a large in-plane upper critical eld
compared with the Pauli paramagnetic eld for a two-
dimensional SC. For example, the Type I Ising super-
conductivity in monolayer MoS 2[48, 49] and NbSe 2[50]
and the Type II Ising superconductivity in monolayer
stanene [51]. Therefore, the interplay of atomic SOC
and the multi-orbital pairing could potentially give rise
to exciting physics. However, to the best of our knowl-
edge, the in
uence of the atomic SOC on the orbital-
dependent pairings remains unsolved. Furthermore, the
multi-orbital nature also gives rise to possible orbital hy-
bridization eects and provides an experimentally con-
trollable handle using lattice strains, both of which could
lead to orbital anisotropy and could potentially change
the pairing symmetry. In particular, lattice strain has
been a useful experimental tool to study unconventional
superconductors [52{54] and has even been proposed to
induce the elusive charge-4e phase [55]. We will be do-
ing an extensive investigation on all the aforementioned
multi-orbital eects.
Another topic of this work is concerned with the co-
existence of TRB pairings and spin magnetism even in
a spin-singlet SC. It is well-known that spin-polarizationarXiv:2201.06265v2 [cond-mat.supr-con] 1 Mar 20232
(SP) can be generated by nonunitary spin-triplet super-
conductivity, which is believed to be the case for LaNiC 2
[56] and LaNiGa 2[57, 58]. More recently, the coexis-
tence of magnetism and spin-singlet superconductivity
is experimentally suggested in multi-orbital SCs, such
as iron-based superconductors [59, 60] and LaPt 3P [61].
Therefore, in addition to the spin-triplet theory, it will be
interesting to examine how SP develops in multi-orbital
spin-singlet SCs as spontaneous TRS breaking in the ab-
sence of external magnetic elds or Zeeman elds.
In this work, we address the above two major issues by
studying a two-band SC with two atomic orbitals (e.g.,
dxzanddyz). We start with the construction of a kp
model Hamiltonian on a square lattice with applied lat-
tice strain. The breaking of C4vdown toC2point group
generally leads to the degeneracy lifting of dxzanddyz.
Based on this model, we study the stability of supercon-
ductivity and the realization of 2D topological supercon-
ductors in both class D and DIII. First and foremost, the
in
uence of atomic SOC is studied, which gives birth to
a new spin-orbit-coupled SC. This exotic state shows the
following features: rstly, a large Pauli-limit violation
is found for the orbital-independent pairing part, which
belongs to the Type II Ising superconductivity. Further-
more, the orbital-dependent pairing part also shows a
weak Pauli-limit violation even though it does not belong
to the family of Ising SCs. Secondly, topological super-
conductivity can be realized with a physical set of pa-
rameters even in the absence of external magnetic elds
or Zeeman elds. In addition, a spin-polarized super-
conducting state could be energetically favored with the
spontaneous breaking of time-reversal symmetry. Our
work implies a new mechanism for the establishment of
spin magnetism in the spin-singlet SC. In the end, we also
discuss how to detect this eect by spin-resolved scanning
tunneling microscopy measurements.
The paper is organized as follows: in section II, we
discuss a two-orbital normal-state Hamiltonian on a 2D
square lattice and also its variants caused by applied in-
plane strain eects, then we show the spin-singlet unitary
or non-unitary pairing states with or without TRS. The
strain eect on pairing symmetries is also studied based
on a weak-coupling theory. In section III, the eects of
atomic SOC on such pairing states are extensively stud-
ied, as well as the in-plane paramagnetic depairing eect.
Besides, the topological superconductivity is studied in
section IV even in the absence of external magnetic elds
or Zeeman elds, after which we consider the spontaneous
TRB eects in section V and show that spin-singlet SC-
induced spin magnetism could emerge in the presence of
orbital SOC. In the end, a brief discussion and conclu-
sion are given in section VI. We will also brie
y comment
on a very recent experiment [62], demonstrating that a
fully gapped superconductor becomes a nodal phase by
substituting S into single-layer FeSe/SrTiO 3.II. MODEL HAMILTONIAN
In this section, we rst discuss the normal-state Hamil-
tonian that will be used throughout this work for an
electronic system consisting of both spin and two locally
degenerate atomic orbitals (e.g., dxzanddyz) on a 2D
square lattice. We assume each unit cell contains only one
atom, so there is no sublattice degree of freedom. The
orbital degeneracy can be reduced by applying the in-
plane lattice strain because the original C4vpoint group
is reduced down to its subgroup C2vfor strain10;01or
11(A more generic strain would reduce the symmetry
directly to C2). Heren1n2represents the strain tensor
whose form will be given later. We will apply the sym-
metry analysis to construct the strained Hamiltonian in
the spirit of kptheory. Then, we discuss the pairing
Hamiltonian and the corresponding classication of spin-
singlet pairing symmetries including non-unitary pairing
states. The strain eect is also investigated on the super-
conducting pairing symmetries based on a weak-coupling
scheme [10].
A. Normal-state Hamiltonian
In this subsection, we construct the two-orbital
normal-state Hamiltonian H0(k) with lattice strain-
induced symmetry-breaking terms. Before that, We rst
showH0(k) in the absence of external lattice strains.
For a square lattice as illustrated in Fig. 1 (a), it owns
theC4vpoint group that is generated by two symme-
try operators: a fourfold rotation symmetry around the
^z-axisC4z: (x;y)!(y; x) and a mirror re
ection
about the ^y ^zplaneMx: (x;y)!( x;y). Other
symmetries can be generated by multiplications, such
as the mirror re
ection about the (^ x+ ^y) ^zplane
Mx+y: (x;y)!(y;x) is given by C4zMx. In the
absence of Rashba spin-orbit coupling (SOC), the sys-
tem also harbors inversion symmetry I, enlarging the
symmetry group to D4h=C4v
fE;Ig. In the spirit of
kpexpansion around the point or the Mpoint, we
consider a two-orbital system described by the inversion-
symmetric Hamiltonian in two dimensions (2D),
H0(k) =(k)00+soc23+o[go(k)]0;(1)
where the basis is made of fdxz;dyzg-orbitals y
k=
(cy
dxz;"(k);cy
dxz;#(k);cy
dyz;"(k);cy
dyz;#(k)). Herecyis the
creation operator of electrons, andare Pauli ma-
trices acting on the orbital and spin subspace, respec-
tively, and 0,0are 2-by-2 identity matrices. Besides,
(k) = (k2
x+k2
y)=2m is the band energy measured
relative to the chemical potential ,mis the eective
mass,socis the atomic SOC [63{65] and ocharacter-
izes the strength of orbital hybridization. This model
could describe the two hole pockets of iron-based super-
conductors [66, 67]. Moreover, the rst two components
ofgo(k) are for the inter-orbital hopping term, while the3
𝑀𝑥
(a)
𝐶4𝑧 𝑥𝑦
𝑑𝑥𝑧orbital
++−
−𝑥𝑧𝑑𝑦𝑧orbital
++−
−𝑦𝑧𝑪𝟒𝒗𝑪𝟐𝒗
𝑪𝟐𝒗
𝑪𝟐𝑪𝟐𝒗𝑴𝒙𝑴𝒚 Basis Orbitals
𝐴1++𝑥2,𝑦2
𝐴2−−𝑥𝑦
𝐵1−+𝑥𝑧𝑑𝑥𝑧
𝐵2+−𝑦𝑧𝑑𝑦𝑧Strain𝜎10(b)
𝑀𝑥𝑀𝑦
(c)
Strain𝜎11
(d)𝑪𝟐𝒗𝑴𝒙′𝑴𝒚′ Basis Orbitals
𝐴1++𝑥′2,𝑦′2
𝐴2−−𝑥′𝑦′
𝐵1−+𝑥′𝑧𝑑𝑥′𝑧
𝐵2+−𝑦′𝑧𝑑𝑦′𝑧
Strain𝜎1𝛿𝐶2𝑧
𝐶2𝑧
𝐶2𝑧𝑪𝟐𝑪𝟐𝒛 Basis Orbitals
𝐴1+𝑥2,𝑦2,𝑥𝑦
𝐵1−𝑦𝑧,𝑥𝑧𝑑𝑥𝑧,𝑑𝑦𝑧Strain𝑥𝑦
𝑥𝑦
FIG. 1. The strain eect on a two-dimensional square lattice. In the absence of lattice strain, (a) shows the square lattice
owing theC4vpoint group that is generated by C4zandMx. We consider the normal-state Hamiltonian with dxz;dyz-orbitals.
Inversion symmetry ( I) is broken by growing crystal samples on an insulating substrate. The in-plane strain eects on the
square lattice are illustrated in (b-d) for applied strain along dierent directions. (b) shows that the ^ xor ^y-axis strain breaks
the square lattice into the rectangular lattice with two independent mirror re
ection symmetries MxandMy, obeying the
subgroupC2vofC4v. TheC2vpoint group contains four one-dimensional irreducible representations (irrep.) A1;A2;B1;B2.
(c) shows that the strain along the ^ x+ ^y-direction also reduces the C4vdown toC2v. (d) represents a general case, where the
subgroupC2is preserved that only has A1andB1irreps.
third term is for the anisotropic eective mass, explained
below in detail.
TheC4v(orD4h) point group restricts go(k) =
(aokxky;0;k2
x k2
y), whereao= 2 is a symmetric case
that increases the C4zto a continues rotational sym-
metry about the ^ z-axis. To be precise, the g1-term,
2okxky10, is attributed to the inter-orbital hopping
integral along the ^x^ydirections,
o
2(cy
dxz;(ix;iy)cdyz;(ix+ 1;iy+ 1)
+cy
dxz;(ix;iy)cdyz;(ix 1;iy 1)
cy
dxz;(ix;iy)cdyz;(ix+ 1;iy 1)
cy
dxz;(ix;iy)cdyz;(ix 1;iy+ 1) + h.c.) ;(2)
where (ix;iy) represents the lattice site. In addition, the
g3-term,o(k2
x k2
y)30, causes the anisotropic eective
masses. For example, the eective mass of the dxzorbital
is1
1=m 2oalong the ^x-axis while that is1
1=m+2oalong
the ^y-axis. This means that the hopping integrals aredierent along ^ xand ^ydirections,
(1
2m o)cy
dxz;(ix;iy)cdxz;(ix+ 1;iy)
+(1
2m+o)cy
dxz;(ix;iy)cdxz;(ix;iy+ 1)
+(1
2m+o)cy
dyz;(ix;iy)cdyz;(ix+ 1;iy)
+(1
2m o)cy
dyz;(ix;iy)cdyz;(ix;iy+ 1) + h.c.:(3)
In this work, we focus on a negative eective mass case
by choosing 1 =m2o>0. However, using a posi-
tive eective mass does not change our main conclusion.
Moreover, our results can be generally applied to other
systems with two orbitals px;py, once it satises the C4v
point group.
The time-reversal symmetry operator is presented as
T=i02KwithKbeing the complex conjugate. And
the inversion symmetry is presented as I=00. It is
easy to show Eq. (1) is invariant under both TandI.
However, inversion can be broken by growing the sample
on insulating substrates, the asymmetric Rashba SOC is
described by
HR(k) =R0[gR(k)]; (4)4
whereRis the strength of the Rashba SOC with
gR(k) = ( ky;kx;0) as required by the C4vpoint group.
Next, we consider the lattice strain eect on the two-
dimensional crystal with a square lattice, as summarized
in Fig. 1 (b-d). The in-plane strain eect is characterized
by the 2-by-2 strain tensor whose elements are dened
asij=1
2
@xiuj+@xjui
, whereuiis the displacement
atralong the ^eidirection. Even though it is an abuse of
notation, it should be self-evident that the here does
not represent the Pauli matrices. The strain tensor can
be parametrized as the following
=cos2cossin
cossin sin2
; (5)
whereis the polar angle with respect to the ^ x-axis. For
the= 0 (=2) case, the compressive or tensile strain
applied along the ^ x-axis (^y-axis) makes the square lattice
as a rectangular lattice, as illustrated in Fig. 1 (b). And
the==4 case is for the shear strain along the (^ x+ ^y)-
direction in Fig. 1 (c). All the above cases reduce the
C4vpoint group into its subgroup C2vthat is generated
by two independent mirror re
ections. Otherwise, it is
generally reduced to C2. The irreducible representations
forC2vandC2are shown in Fig. 1 (b-d). Based on the
standard symmetry analysis, to the leading order, the
strained Hamiltonian is given by
Hstr=tstr[sin(2)1+ cos(2)3]0; (6)
where both tstrandcan be controlled in experi-
ments [68]. AndHstrcan be absorbed into the go-vector
in Eq. (1), renormalizing the orbital hybridization as ex-
pected. Furthermore, one can check that Hstrpreserves
bothTandI, but explicitly breaks the C4z=i2ei
43
because of [Hstr;C4z]6= 0. Interestingly, the orbital tex-
ture on the Fermi surface can be engineered by strain,
and its eect on superconducting pairing symmetries is
brie
y discussed in the Appendix C.
Therefore, a strained normal-state Hamiltonian is
HN(k) =H0(k) +HR(k) +Hstr; (7)
which will be used throughout this work. The Rashba
SOC induced spin-splitting bands are considered only
when we discuss the topological superconducting phases
in section IV and V, even though the normal-state Hamil-
tonianHN(k) is topologically trivial. For the supercon-
ducting states, we focus on the inversion symmetric pair-
ings (i.e., spin-singlet s-wave pairing) and their response
to applied strains or in-plane magnetic elds.
In the absence of Rashba SOC, the band structures of
HN(k) in Eq. (7) are given by
E(k) = 1
2m(k2
x+k2
y)q
2soc+ ~g2
1+ ~g2
3;(8)
where we dene the strained orbital hybridization ~gvec-
tor with ~g1=aookxky+tstrsin(2) and ~g3=o(k2
x
k2
y)+tstrcos(2). Each band has two-fold degeneracy, en-
forced by the presence of both TandI. At the point
𝑘𝑥𝑘𝑦
𝑘𝑥𝑘𝑦(a) (b)FIG. 2. The lattice strain eect on the Fermi surfaces of the
normal-state Hamiltonian without Rashba SOC. (a) shows
the two Fermi surfaces without lattice strain (i.e., tstr= 0),
thusC4z-symmetric energy contours are formed. (b) shows
the breaking of C4zby lattice strain with tstr= 0:4 and
= 0, onlyC2z-symmetric energy contours appear. Other
parameters used here are m= 0:5;a0= 1;o= 0:4,R= 0
and= 0:5.
(kx=ky= 0),E
=p
2soc+t2
str. The two Fermi sur-
faces with and without strain are numerically calculated
and shown in Fig. 2, where we choose < |