Weyl spin-orbit-coupling-induced interactions in uniform and trapped atomic quantum fluids

12
PHYSICAL REVIEW A 88, 053607 (2013) Weyl spin-orbit-coupling-induced interactions in uniform and trapped atomic quantum fluids Reena Gupta, 1 G. S. Singh, 1,* and J¨ urgen Bosse 2 1 Physics Department, Indian Institute of Technology Roorkee, Roorkee 247 667, India 2 Institute of Theoretical Physics, Freie Universit¨ at, D-Berlin 14195, Germany (Received 23 July 2013; published 7 November 2013) We establish through analytical and numerical studies of thermodynamic quantities for noninteracting atomic gases that the isotropic three-dimensional spin-orbit coupling, the Weyl coupling, induces interaction which counters “effective” attraction (repulsion) of the exchange symmetry present in zero-coupling Bose (Fermi) gas. The exact analytical expressions for the grand potential and hence for several thermodynamic quantities have been obtained for this purpose in both uniform and trapped cases. It is enunciated that many interesting features of spin-orbit-coupled systems revealed theoretically can be understood in terms of coupling-induced modifications in statistical interparticle potential. The temperature dependence of the chemical potential, specific heat, and isothermal compressibility for a uniform Bose gas is found to have signature of the incipient Bose-Einstein condensation in the very weak coupling regime although the system does not really go in the Bose-condensed phase. The transition temperature in the harmonically trapped case decreases with an increase of coupling strength consistent with the weakening of the statistical attractive interaction. Anomalous behavior of some thermodynamic quantities, partly akin to that in dimensions less than two, appears for uniform fermions as soon as the Fermi level goes down the Dirac point on increasing the coupling strength. It is suggested that the fluctuation-dissipation theorem can be utilized to verify anomalous behaviors from studies of long-wavelength fluctuations in bunching and antibunching effects. DOI: 10.1103/PhysRevA.88.053607 PACS number(s): 03.75.Ss, 05.30.Fk, 05.30.Jp, 67.85.Lm I. INTRODUCTION It has been during the last few years that theoretical proposals came up to have synthetic spin-orbit coupling (SOC) in neutral atoms resulting from laser-induced coupling between the internal psuedospin- 1 2 degree of freedom and the atomic center-of-mass motion (see [1,2] and references therein). The required process engineers a spatially homo- geneous but non-Abelian vector potential and yields a term in the Hamiltonian which has the form of Rashba or linear Dresselhaus SO coupling. Both these two-dimensional (2D) couplings, appearing in a natural way in solids under separate physical conditions, have been known for a long time [3] but are being difficult to implement in atomic systems. However, equal Rashba-Dresselhaus, which is one-dimensional (1D) and is now also called [4] the NIST SOC due to its first implementation by Spielman’s group at the NIST [5], has been realized experimentally in bosonic [5,6] as well as fermionic [7,8] atomic gases. Recently, methods to generate three-dimensional (3D) analogues of the Rashba type have also been suggested [9,10]. The proposed 3D couplings for cold atoms could be anisotropic or rotationally invariant. The latter case has been christened by Anderson et al. [9] as the Weyl SOC which is, in fact, an aide m´ emoire to the Weyl fermions [11] that obey the Dirac-like equation [12]. Theoretical studies have been accomplished for a variety of physical phenomena (see references in reviews [1,13,14]), considering variations in the Zeeman field, trapping potential, optical-lattice parameters, detuning, and s -wave scattering length a s apart from the SO coupling strength. The significant predictions for bosonic systems are that the SOC makes them behave in ways that have no known analogues in conventional * [email protected] condensed-matter systems, namely the half-vortex phase [15] and the striped superfluid phase [16]. Also, unconventional BEC beyond the “no-node” theorem [17] has been analyzed in [13,18] and Zitterbewegung oscillations have been observed experimentally [19]. Some of the salient features which have emerged in studies of fermions are nontrivial topological order [20,21], enhanced pairing field [9,20,2224] and exotic pairing states [4], possible existence of Majorana fermions [25] and the Fulde-Ferrell-Larkin-Ovchinnikov phase [4,26], and existence of Dirac cones in honeycomb optical lattices [27]. These aspects have provided additional impetus to the study of ultracold atoms and matter waves which have been at the forefront for the last two decades to mimic many physical phenomena [28]. It may be recalled in the context of many-body aspects that quantum gases have statistical interactions meaning thereby that ideal bosons (fermions) have “effective” interparticle attraction (repulsion) as compared to a Boltzmann gas as a consequence of their total wave function being symmetric (antisymmetric) under exchange of a pair of particles (see, for example, Ref. [29]). Thus the two kinds of quantum particles which have either integral or half-odd-integral spin had earlier been considered to be distinct [30] both at the single-particle level and the many-particle level. However, it is noteworthy in the context of artificial SO coupling that pseudospin- 1 2 bosons have not only been conceptualized but also experimentally realized. Although the concept is not in conflict with the Pauli spin-statistics theorem [31] because the symmetry operations in the pseudospin space are not related to real-space rotations, it implies that bosons and fermions could be made identical at the single-particle level. Despite availability of enormous literature involving all kinds of complexities in the field, a basic question of how the SOC modifies the many-body effect mentioned above is not known. The main purpose of this paper is to address this aspect both in uniform and harmonically 053607-1 1050-2947/2013/88(5)/053607(12) ©2013 American Physical Society

Transcript of Weyl spin-orbit-coupling-induced interactions in uniform and trapped atomic quantum fluids

PHYSICAL REVIEW A 88, 053607 (2013)

Weyl spin-orbit-coupling-induced interactions in uniform and trapped atomic quantum fluids

Reena Gupta,1 G. S. Singh,1,* and Jurgen Bosse2

1Physics Department, Indian Institute of Technology Roorkee, Roorkee 247 667, India2Institute of Theoretical Physics, Freie Universitat, D-Berlin 14195, Germany

(Received 23 July 2013; published 7 November 2013)

We establish through analytical and numerical studies of thermodynamic quantities for noninteracting atomicgases that the isotropic three-dimensional spin-orbit coupling, the Weyl coupling, induces interaction whichcounters “effective” attraction (repulsion) of the exchange symmetry present in zero-coupling Bose (Fermi) gas.The exact analytical expressions for the grand potential and hence for several thermodynamic quantities havebeen obtained for this purpose in both uniform and trapped cases. It is enunciated that many interesting features ofspin-orbit-coupled systems revealed theoretically can be understood in terms of coupling-induced modificationsin statistical interparticle potential. The temperature dependence of the chemical potential, specific heat, andisothermal compressibility for a uniform Bose gas is found to have signature of the incipient Bose-Einsteincondensation in the very weak coupling regime although the system does not really go in the Bose-condensedphase. The transition temperature in the harmonically trapped case decreases with an increase of couplingstrength consistent with the weakening of the statistical attractive interaction. Anomalous behavior of somethermodynamic quantities, partly akin to that in dimensions less than two, appears for uniform fermions assoon as the Fermi level goes down the Dirac point on increasing the coupling strength. It is suggested that thefluctuation-dissipation theorem can be utilized to verify anomalous behaviors from studies of long-wavelengthfluctuations in bunching and antibunching effects.

DOI: 10.1103/PhysRevA.88.053607 PACS number(s): 03.75.Ss, 05.30.Fk, 05.30.Jp, 67.85.Lm

I. INTRODUCTION

It has been during the last few years that theoreticalproposals came up to have synthetic spin-orbit coupling(SOC) in neutral atoms resulting from laser-induced couplingbetween the internal psuedospin- 1

2 degree of freedom andthe atomic center-of-mass motion (see [1,2] and referencestherein). The required process engineers a spatially homo-geneous but non-Abelian vector potential and yields a termin the Hamiltonian which has the form of Rashba or linearDresselhaus SO coupling. Both these two-dimensional (2D)couplings, appearing in a natural way in solids under separatephysical conditions, have been known for a long time [3] butare being difficult to implement in atomic systems. However,equal Rashba-Dresselhaus, which is one-dimensional (1D)and is now also called [4] the NIST SOC due to its firstimplementation by Spielman’s group at the NIST [5], hasbeen realized experimentally in bosonic [5,6] as well asfermionic [7,8] atomic gases. Recently, methods to generatethree-dimensional (3D) analogues of the Rashba type havealso been suggested [9,10]. The proposed 3D couplings forcold atoms could be anisotropic or rotationally invariant. Thelatter case has been christened by Anderson et al. [9] as theWeyl SOC which is, in fact, an aide memoire to the Weylfermions [11] that obey the Dirac-like equation [12].

Theoretical studies have been accomplished for a varietyof physical phenomena (see references in reviews [1,13,14]),considering variations in the Zeeman field, trapping potential,optical-lattice parameters, detuning, and s-wave scatteringlength as apart from the SO coupling strength. The significantpredictions for bosonic systems are that the SOC makes thembehave in ways that have no known analogues in conventional

*[email protected]

condensed-matter systems, namely the half-vortex phase [15]and the striped superfluid phase [16]. Also, unconventionalBEC beyond the “no-node” theorem [17] has been analyzedin [13,18] and Zitterbewegung oscillations have been observedexperimentally [19]. Some of the salient features which haveemerged in studies of fermions are nontrivial topologicalorder [20,21], enhanced pairing field [9,20,22–24] and exoticpairing states [4], possible existence of Majorana fermions [25]and the Fulde-Ferrell-Larkin-Ovchinnikov phase [4,26], andexistence of Dirac cones in honeycomb optical lattices [27].These aspects have provided additional impetus to the studyof ultracold atoms and matter waves which have been at theforefront for the last two decades to mimic many physicalphenomena [28].

It may be recalled in the context of many-body aspects thatquantum gases have statistical interactions meaning therebythat ideal bosons (fermions) have “effective” interparticleattraction (repulsion) as compared to a Boltzmann gas as aconsequence of their total wave function being symmetric(antisymmetric) under exchange of a pair of particles (see, forexample, Ref. [29]). Thus the two kinds of quantum particleswhich have either integral or half-odd-integral spin had earlierbeen considered to be distinct [30] both at the single-particlelevel and the many-particle level. However, it is noteworthy inthe context of artificial SO coupling that pseudospin- 1

2 bosonshave not only been conceptualized but also experimentallyrealized. Although the concept is not in conflict with the Paulispin-statistics theorem [31] because the symmetry operationsin the pseudospin space are not related to real-space rotations,it implies that bosons and fermions could be made identicalat the single-particle level. Despite availability of enormousliterature involving all kinds of complexities in the field, abasic question of how the SOC modifies the many-body effectmentioned above is not known. The main purpose of this paperis to address this aspect both in uniform and harmonically

053607-11050-2947/2013/88(5)/053607(12) ©2013 American Physical Society

REENA GUPTA, G. S. SINGH, AND JURGEN BOSSE PHYSICAL REVIEW A 88, 053607 (2013)

trapped gases. The closer analysis of interesting features foundin some of the theoretical studies enumerated below suggeststhat the SO coupling weakens the statistical interactions inbosons as well as fermions and we establish it through studyof the thermodynamic properties.

It is a textbook result for noninteracting bosons that thereis a Bose-condensed phase at T < TC for the 3D gas, atT = 0 for the 2D gas, and no condensation at all in the1D gas. On the other hand, it has recently been pointed outfor the Rashba-coupled noninteracting gas that there is nocondensation at all in the 2D case [32] and TC goes down tozero in the 3D case [33,34] consequent upon the single-particledensity of states (DOS) acquiring in the low-energy regionbehavior of reduced spatial dimensionality. The completedestruction of 3D condensate by the Weyl coupling at allT , including at T = 0, in the noninteracting gas has beenpointed out in Refs. [32,35] to be due to appearance ofcoupling-weighted 1D DOS. The enhanced single-particleDOS at the Fermi surface due to SOC has been utilizedto understand also that the two-body bound state appearsin an SO-coupled Fermi gas on the BCS side (as < 0) ofthe Feshbach resonance howsoever weak the attraction mightbe [9,22] and that the interaction effect gets magnified resultingin increased fermionic-superfluid transition temperature [24].In addition, there is BCS-BEC crossover on increasing SOCwhile keeping as fixed even for as → 0− [36]. Cui andZhou [32] and Zhou and Cui [37] have studied the effect ofSOC on an interacting bosonic system and have found thatboth Rashba and 3D couplings create condensate depletion.Zheng et al. [38] have discussed a shift in the BEC transitiontemperature in an SO-coupled interacting Bose gas. Han andSa de Melo [39] have arrived at the conclusion that the Rashbamakes the Pauli pressure to decrease with an increase in thecoupling strength even if as is kept fixed. Gong et al. [20]have shown that the fermionic superfluidity destroyed by theZeeman field above a critical value can be restored by a finiteSOC without changing as . Yu [40] has noticed that the SOCenhances the short-distance correlation length in Fermi gases.

Exact analytical expressions for grand potentials of WeylSO-coupled noninteracting systems in uniform as well astrapped systems have been obtained in unified forms introduc-ing a parameter η to represent Bose (η = +1), Fermi (η = −1),and Boltzmann (η = 0) gases. The analytical expressionsfollowing therefrom for various thermodynamic quantitiesof the uniform system have universal character for a fixeddensity, when ε0

F serves as an energy scale, as they dependultimately only on two dimensionless physical parameters:reduced temperature T = T/T 0

F and reduced SO-couplingstrength κ = κ/k0

F with ε0F, T 0

F and k0F being, respectively, the

Fermi energy, Fermi temperature, and Fermi wave vector foran ideal gas without SOC. The graphical depiction of thepressure as a function of SOC strength for fixed temperatureand density shows that “effective” interactions become lessattractive (repulsive) in uniform as well as trapped bosons(fermions), a result which is corroborated by plots of theisothermal compressibility and analytically in uniform gasesby the virial expansion of the equation of state (EOS).

It may be noted that some of the thermodynamic propertiesof SO-coupled systems had been studied earlier by He andYu [35], who found a few interesting results. But their studies

were restricted to the uniform gases, and also their mainfocus was in high- and low-temperature regimes whereinanalytical expressions could be obtained. Our studies coveran entire range of physical parameters not only for uniformgases but for the trapped systems as well and we retrievetheir results as limiting cases of our expressions. In addition,we provide a sound theoretical basis for (i) the observationby Han and Sa de Melo [39], while studying theoreticallyBCS-BEC crossover in a balanced mixture of 3D ultracoldfermions, that the Pauli pressure decreases and consequentlythe isothermal compressibility increases on increase in theRashba coupling strength, and (ii) the finding of Ghosh et al.[41] regarding decrease in Pauli pressure with increase of 3DSOC in uniform systems while studying trapped fermions.Furthermore, we find signature of incipient Bose condensationfor very weak SO coupling in a uniform gas although thereis really no condensation due to the presence of coupling-weighted 1D DOS. We observe anomalous behavior, typicalof spatial-dimensional crossover, in some thermodynamicquantities of uniform fermions when the Fermi level crossesthe Dirac point on varying the coupling strength. Of course,the physics of harmonically trapped gases is more interestingas the phenomenon of BEC destroyed by the SOC in 3D Bosegas is restored by trapping even in the noninteracting case;our expressions offer opportunity for the study of completethermodynamics of such systems.

The paper is structured as follows. Section II starts with anoverview of basic aspects pertaining to the Weyl SO-coupledHamiltonian for noninteracting atomic quantum gases, and thebehavior of the resulting DOS is described for both uniformand trapped gases. The analytical expressions for grand po-tentials involving polylogarithmic functions are presented forthese systems in Secs. III A and III B, respectively, and exactanalytical expressions for various thermodynamic quantitiesare developed in both the cases. Computed results are depictedgraphically in Sec. IV wherein detailed discussions are alsogiven. Section V gives summary and outlook of our work.Appendix A discusses some aspects pertaining to plots of DOSin uniform gases, Appendix B contains derivation for the ex-pression of the isochoric specific heat, and detailed discussionof the fermionic chemical potential is given in Appendix C.

II. WEYL SPIN-ORBIT-COUPLED SYSTEMS

A. Basic aspects

We consider a system of cold neutral atoms, bosons orfermions, having multilevel hyperfine states and moving in thepresence of spatially varying laser fields. The coupling inducedvia two-photon Raman transitions linking four internal levelsin a tetrahedral topology, as described in [9], gives rise to aneffective non-Abelian vector potential A = −hκσ , where κ

is the isotropic coupling strength and σ = (σx,σy,σz) is theconventional Pauli spin operator in the pseudospin space.Hence the single-particle effective Hamiltonian of (↑, ↓)pseudospin- 1

2 particles would become H = (1/2m)(pI − A)2

which one may take, without any loss of generality, in the form,

H = 1

2m(p2 + h2κ2)I + vσ · p. (1)

053607-2

WEYL SPIN-ORBIT-COUPLING-INDUCED INTERACTIONS . . . PHYSICAL REVIEW A 88, 053607 (2013)

Here v = hκ/m is the coupling strength having unit ofvelocity, I is the 2 × 2 identity matrix, and p is the momentumoperator.

It may be noted that the atoms have acquired a psuedospin- 12

degree of freedom, irrespective of the system without couplingbeing fermionic or bosonic, which has coupled with the atomiccenter-of-mass motion. If the inconsequential constant term isnot retained in Eq. (1), the resulting isotropic coupling termHSO = vσ · p is a natural symmetric extension of the Rashbacoupling to 3D and has the form of the Weyl Hamiltonian[12]. One has thus the Weyl SO coupling, the terminologycoined in [9]. Furthermore, the Hamiltonian H possessescomplete rotational symmetry, the highest such symmetryamong all types of SO couplings, as a consequence of thecommutation relation [J,H] = 0, where J = L + 1

2hσ is thetotal angular momentum operator with L as the atomic center-of-mass orbital angular momentum operator. In addition, theHamiltonian respects the time-reversal symmetry even whenκ �= 0 but the Galilean invariance and the parity conservationare broken. Various aspects of the simplicity inherently presentin the Weyl-coupled Hamiltonian for isotropically trappedgases described by Anderson and Clark [42] apply here inthe case of uniform gases, too.

The eigenvalues ±1 of the helicity operator σ · p/|p| aregood quantum numbers and the Hamiltonian in Eq. (1) can bediagonalized yielding eigenenergies,

ε±(k) = h2

2m(k ± κ)2, (2)

which correspond to two orthogonal eigenstates |k+〉 =uk|k↑〉 + vke

iφk |k↓〉 and |k−〉 = uk|k↓〉 − vke−iφk |k↑〉 with

uk = [(1 + kz/|k|)/2]1/2, vk = [(1 − kz/|k|)/2]1/2, and φk =arg(kx + iky). The system has spherical manifold of zero-energy ground states and there is a 3D Dirac point at k = 0which is robust since the presence of all the three spin matricesin HSO ensures that a homogeneous Zeeman field cannot liftup the degeneracy at this point. The point is located in energyspace at ε+ = ε− = h2κ2/2m which is the spin-orbit-couplingenergy εSO. Its significance gets highlighted below whenwe discuss SO-induced dimensional reduction, and againin Sec. IV when we discuss the appearance of anomalousbehavior in some thermodynamic quantities of fermions assoon as the Fermi level goes down this point on increasingcoupling strength.

Equation (2) implies that the system possesses two constantenergy surfaces (CESs) in k space, both being spherical havingradii k± = k0 ∓ κ , where k0 = (2mε0/h

2)1/2 with ε0 = ε+ =ε− having a chosen constant value. For κ = 0, two spheres havethe same radii (k+ = k− = k0) but as soon as the SO couplingis switched on, we have two concentric spheres with k+ <

k0 and k− > k0, i.e., the volume of the CES correspondingto the positive helicity branch gets decreased whereas thatof the negative branch gets increased. When κ would increase,the smaller(larger) volume CES would shrink(expand) and forκ = k0, the plus branch CES would reduce to a point. Thissequence of interesting behavior together with that for κ >

k0 unveils for fermions the so-called Fermi-surface-topologytransition described in [36,43].

The Hamiltonian (1) can be generalized for a trapped gasto have the semiclassical form [33,42] wherein the trapping

potential U (r) would simply get added on the right-hand sidein Eq. (1) and the counterpart of Eq. (2) would become thesemiclassical energy spectrum given by

ε±(k,r) = h2

2m(k ± κ)2 + U (r), (3)

in the six-dimensional phase space (k,r). The motion of theparticle would now get restricted to the region of space whereinU (r) is less than the particle’s energy ε. This introduces theconcept of the effective volume as described in [44] and canbe expressed for the 3D systems as

V ∗3D(ε) =

∫�(ε − U (r))dr = 4π

3r3

max, (4)

where �(x) is the Heaviside step function and rmax is thesolution of the equation U (rmax) = ε.

B. Density of states

If the energy-level separation is much smaller than thethermal energy kBT , which is usually the situation underexperimental conditions for cold atoms, the eigenenergiescan be taken to be quasicontinuous and the Thomas-Fermisemiclassical approximation would become applicable. Con-sequently, the density of states (DOS) for each branch can becalculated using the expression,

G±(ε) = 1

(2π )3

∫∫d3rd3kδ(ε − ε±(k,r)). (5)

Here δ(x) is the Dirac delta function and d3rd3k/(2π )3 isthe number of quantum states in the phase-space elementaryvolume d3rd3k with (2π )3 being the volume occupied by asingle quantum state.

In order to evaluate the integrals appearing in Eq. (5), werequire the explicit form for U (r). We choose this isotropictrapping potential to have the generic power-law behavior:U (r) = U0(r/0)γ with U0 and 0 as constants. The value ofthe exponent γ defines the nature of the potential; γ = 2 givesthe important case of the harmonic trapping and limγ→∞ U (r)gives the spherical potential well of the radius 0: U (r) = 0for r < 0 and U (r) = ∞ for r > 0. Thus the uniformgas corresponds to the limiting situation when γ → ∞and 0 → ∞.

The evaluation requiring the use of the integral [45],

∫ π/2

0sinμ−1 θ cosν−1 θdθ = 1

2

�(

μ

2

)�

(ν2

)�

2 + ν2

) ;

(6)Reμ > 0, Reν > 0,

has ultimately yielded the expression for the total DOS, G(ε) ≡G(ε,κ,γ ) = G+(ε) + G−(ε), in the form,

G(ε) =(

4π30

3

)(ε

U0

) 3γ

(3

γ+ 1

)

×[

�(

32

)ρ0

3D(ε)

�(

+ 32

) + κ2

�(

12

)ρ0

1D(ε)

�(

+ 12

)]

, (7)

053607-3

REENA GUPTA, G. S. SINGH, AND JURGEN BOSSE PHYSICAL REVIEW A 88, 053607 (2013)

where

ρ0d (ε) = 1

2d−1πd/2�(d/2)

(2m

h2

)d/2

εd/2−1�(ε) (8)

is the DOS per unit volume for both spin orientations of a d-dimensional uniform system of spin- 1

2 particles without SOC,and �(ν) is the gamma function. Thus we have obtained a quitegeneral expression in a neat analytical form for 3D systemsand we now examine its validity in known special cases.

The expression, as given in [46], for the DOS of idealquantum gases under the influence of a power-law externalpotential but without SOC is recovered in the 3D case onsubstituting κ = 0 and 0 = 1 in Eq. (7). In addition, forthe most important case of the harmonic trapping potential,U (r) = 1

2mω2r2, we get

GH (ε) = ε2

h3ω3

(1 + 4

εSO

ε

)�(ε), (9)

whose comparison with the well-known result for harmon-ically trapped Bose gas without SOC reveals the presenceof an extra factor 2 here because we have considered thepseudospin- 1

2 bosons. In order to get the result for SO-coupleduniform gases, however, we first obtain from Eq. (7) theDOS per unit effective volume and then take the limit γ →∞. Since rmax = 0(ε/U0)1/γ , the procedure gives the DOSρ(ε) ≡ ρ3D(ε,κ) as

ρ(ε) = ρ03D(ε)

(1 + εSO

ε

), (10)

a rewritten form of that given in Ref. [35]. The detailedbehavior of this DOS is analyzed in Appendix A wherein it isfound that ρ(ε) possesses a minimum at ε = εSO with εSO =h2κ2/2m being the SO coupling energy which corresponds tothe location of the Dirac point in a uniform gas. The presenceof SOC-weighted 1D DOS in Eq. (10) imparts the systemcharacteristics of reduced spatial dimensionality: quasi-1D(more than 1D) for ε < εSO and quasi-3D (less than 3D)for ε > εSO. Thus the uniform system undergoes dimensionalreduction for large κ analogous to that mentioned in Ref. [42]for SO-coupled harmonically trapped gases.

We have the inequality ρ(ε) � ρ03D(ε) wherein the equality

holds for ε = ∞ or κ = 0. This SOC-induced enhancementin DOS has been utilized in the literature (see, for example,in [9,32,35,36]) to give plausible reasoning to understand someof the interesting features enumerated in the Introduction.However, we assert that not only the explanations for thesefeatures but also for the additional ones mentioned in theIntroduction could be constructed in terms of SOC-inducedmodifications in the statistical interactions amongst particlesof the system described in Sec. IV.

III. ANALYTICAL THERMODYNAMICS

It could be noted that the process of evaluation forthe thermodynamic quantities involves energy integrationand hence once the integration has been accomplished, theexpressions so obtained cannot be used to get results forthe uniform case even if the generic power-law potential isused. Hence we consider here separately both uniform and

harmonically trapped gases having Weyl SOC to obtain exactanalytical expressions for various thermodynamic quantities.

A. Uniform gases

The grand thermodynamic potential �(μ,T ,V ) of Bose(η = +1), Fermi (η = −1), and Boltzmann (η = 0) gases canbe written in a unified form as

� = 1

ηβ

∑k,λ=±

ln[1 − ηz exp{−βελ(k)}]

= V

ηβ

∫ ∞

0ρ(ε) ln(1 − ηze−βε)dε, (11)

where z = eβμ is the fugacity, μ the chemical potential,β = 1/(kBT ) the inverse temperature, and kB the Boltzmannconstant. Using Eq. (10) in Eq. (11), substituting x = βε

and introducing the thermal de Broglie wavelength � =(2πh2β/m)1/2, one gets

� = 4V√πηβ�3

∫ ∞

0

(√x + κ2�2

4π√

x

)ln(1 − ηze−x)dx.

(12)

Now we use the identity,

u

ex − u= d

dxln(1 − ue−x), (13)

in the integral representation [47],

Liν(u) = u

�(ν)

∫ ∞

0

xν−1

ex − udx, (14)

for the polylogarithm of νth order with ν > 0, and then performthe integration by parts. The process yields finally

Liν+2(u) = − 1

�(ν + 1)

∫ ∞

0xν ln(1 − ue−x)dx, (15)

with ν > −1. Hence Eq. (12) is rewritten in a neat analyticalform as

� = − 2V

ηβ�3[Li5/2(ηz) + φ(κ�)Li3/2(ηz)], (16)

with φ(κ�) = κ2�2/2π , and can be utilized to discusscomplete thermodynamics of the system.

The total number of particles and the entropy given by N =−(∂�/∂μ)T ,V and S = −(∂�/∂T )V,μ yield the expressions,

N = 2V

η�3[Li3/2(ηz) + φ(κ�)Li1/2(ηz)], (17)

and

S

kB

= V

η�3[5Li5/2(ηz) + 3φ(κ�)Li3/2(ηz)] − Nβμ, (18)

where we have used the recurrence relation,

∂zLiν(z) = 1

zLiν−1(z), (19)

053607-4

WEYL SPIN-ORBIT-COUPLING-INDUCED INTERACTIONS . . . PHYSICAL REVIEW A 88, 053607 (2013)

obtained from Eq. (14). The relation � = −PV givespressure,

P = 2

ηβ�3[Li5/2(ηz) + φ(κ�)Li3/2(ηz)]. (20)

The internal energy and the Helmholtz free energy can beobtained through Legendre transformations E = � + T S +μN and F = � + μN . Thus,

E = V

ηβ�3[3Li5/2(ηz) + φ(κ�)Li3/2(ηz)], (21)

and

F = μN − 2V

ηβ�3[Li5/2(ηz) + φ(κ�)Li3/2(ηz)], (22)

and the EOS is

PV = 2

3

[E +

(∂E

∂ ln κ

], (23)

a form obtained also in Ref. [35]. On using Eq. (17) in thedefinition KT = −(1/V )(∂V/∂P )T ,N = (1/n)(∂n/∂P )T ,N =(1/n2)(∂n/∂μ)T ,N , the isothermal compressibility takes the

form,

KT = 2β

ηn2�3[Li1/2(ηz) + φ(κ�)Li−1/2(ηz)]. (24)

For |u| < 1, Eq. (14) has the series expansion Liν(u) =∑∞=1(u/ν). Hence for |ηz| < 1, Eqs. (17), (21), and (20)

give

N = 2V

η�3

∞∑=1

(ηz)(

1

3/2+ φ(κ�)

1

1/2

), (25)

E = V

ηβ�3

∞∑=1

(ηz)(

3

5/2+ φ(κ�)

1

3/2

), (26)

and

P = 2

ηβ�3

∞∑=1

(ηz)(

1

5/2+ φ(κ�)

1

3/2

). (27)

For bosons and fermions, the above three equations yield,respectively, Eqs. (10), (9), and (8) of Ref. [35] valid in thehigh-temperature limit when fugacity z < 1. Additionally, ifone considers the Boltzmann (η = 0) gas, only the leadingterm, i.e., = 1, survives in the expression for N , E or P ontaking the limit η → 0.

The evaluation for the specific heat CV = (∂E/∂T )S,V hasbeen presented in Appendix B and the final expression is

CV

NkB

= 3

4

[5Li5/2(ηz) + φ(κ�)Li3/2(ηz)]

[Li3/2(ηz) + φ(κ�)Li1/2(ηz)]− 1

4

[3Li3/2(ηz) + φ(κ�)Li1/2(ηz)][3Li3/2(ηz) + φ(κ�)Li1/2(ηz)]

[Li3/2(ηz) + φ(κ�)Li1/2(ηz)][Li1/2(ηz) + φ(κ�)Li−1/2(ηz)]. (28)

The evaluation for thermal expansion coefficient αT = (1/V )(∂V/∂T )P,N = KT(∂P/∂T )V,N involves some aspects as in thecase of the evaluation of CV , and we obtain finally

T αT = 5

2

[Li5/2(ηz) + φ(κ�)Li3/2(ηz)][Li1/2(ηz) + φ(κ�)Li−1/2(ηz)]

[Li3/2(ηz) + φ(κ�)Li1/2(ηz)]2− φ(κ�)Li3/2(ηz)

[Li1/2(ηz) + φ(κ�)Li−1/2(ηz)]

[Li3/2(ηz) + φ(κ�)Li1/2(ηz)]2

−1

2

[3Li3/2(ηz) + φ(κ�)Li1/2(ηz)]

[Li3/2(ηz) + φ(κ�)Li1/2(ηz)]. (29)

It can be noted that on taking the limit κ → 0, all ourexpressions with the SOC reduce to the corresponding resultsfor ideal quantum gases in terms of polylogs [48] and tothose given in textbooks (see, for example, [49]) in termsof Bose and Fermi functions. In addition, the thermodynamicrelations CP = CV + V T α2

T/KT and KS = KT − V T α2T/CP,

where KS is the adiabatic compressibility, suggest that thestudy of complete thermodynamics is possible with the helpof the expressions derived above.

B. Harmonic trapping

The harmonically trapped thermodynamic systems arecharacterized by the extensive variable, “harmonic” volumeV = 1/ω3, and the conjugate intensive variable, “harmonic”pressure P = −(∂�H/∂V)μ,T . The semiclassical expressionfor the grand potential is obtained using Eq. (9) togetherwith Eq. (14) in �H (μ,T ,V) = (1/ηβ)

∫ ∞0 GH (ε) ln(1 −

ηze−βε)dε and is given by

�H = − 2Vηh3β4

[Li4(ηz) + φ(κ�)Li3(ηz)]. (30)

The number of particles and the energy are now given by

N = 2Vηh3β3

[Li3(ηz) + φ(κ�)Li2(ηz)], (31)

E = 2Vηh3β4

[3Li4(ηz) + 2φ(κ�)Li3(ηz)], (32)

and the EOS can be written in the form,

PV = 1

3

[E + 1

4

(∂E

∂ ln κ

]. (33)

This gives PV = 13E for κ → 0 as the EOS of a harmonically

trapped gas without SOC is just like that of ultrarelativistic ormassless particles [50].

The expressions for the other quantities are now

P = 2

ηh3β4[Li4(ηz) + φ(κ�)Li3(ηz)], (34)

F = μN − 2Vηh3β4

[Li4(ηz) + 2φ(κ�)Li3(ηz)], (35)

S

NkB

= 4Li4(ηz) + 3φ(κ�)Li3(ηz)

Li3(ηz) + φ(κ�)Li2(ηz)− βμ, (36)

053607-5

REENA GUPTA, G. S. SINGH, AND JURGEN BOSSE PHYSICAL REVIEW A 88, 053607 (2013)

KT =(

β

n

)Li2(ηz) + φ(κ�)Li1(ηz)

Li3(ηz) + φ(κ�)Li2(ηz), (37)

CVNkB

= 6[2Li4(ηz) + φ(κ�)Li3(ηz)]

Li3(ηz) + φ(κ�)Li2(ηz)− [3Li3(ηz) + 2φ(κ�)Li2(ηz)]2

[Li3(ηz) + φ(κ�)Li2(ηz)][Li2(ηz) + φ(κ�)Li1(ηz)], (38)

and

T αT = 4[Li4(ηz) + φ(κ�)Li3(ηz)][Li2(ηz) + φ(κ�)Li1(ηz)]

[Li3(ηz) + φ(κ�)Li2(ηz)]2− φ(κ�)Li3(ηz)

[Li2(ηz) + φ(κ�)Li1(ηz)]

[Li3(ηz) + φ(κ�)Li2(ηz)]2

− [3Li3(ηz) + 2φ(κ�)Li2(ηz)]

[Li3(ηz) + φ(κ�)Li2(ηz)]. (39)

All the expressions in this subsection are new and give,under the limit κ → 0, the results for the harmonically trappedideal gases without SOC; the corresponding Bose gas results[50] are thus retrieved keeping in mind that here we are dealingwith pseudospin- 1

2 bosons. Furthermore, the thermodynamicrelations involving CP ,CV ,αT,KT, and KS mentioned in theprevious subsection become helpful here, too.

IV. NUMERICAL RESULTS AND DISCUSSIONS

The variation of rescaled pressure P = P/P0 with T (or κ)for various fixed values of κ (or T ) is shown in Figs. 1(a)–1(d).Here P0 = (2nε0

F/5) is the pressure of spin- 12 fermions without

SOC at T = 0 and n is the number density. It can be seenfrom plots that pressure decreases for fermions but increasesfor bosons with increase in the Weyl coupling at any fixedtemperature. In order to have a closer look at this behavioranalytically, we proceed to construct the virial expansion forthe EOS in the form,

Pβ =∞∑

=1

b(T ,κ)n, (40)

Fermions a

0.00.5

1.03.0

0.0 0.1 0.2 0.3 0.4 0.50.0

0.5

1.0

1.5

T TF0

PP 0

Fermions b 0.1

0.2

0.3

0.4

0.0 0.2 0.4 0.6 0.8 1.00.4

0.6

0.8

1.0

1.2

1.4

1.6

Κ kF0

PP 0

Bosonsc

0.00.5

3.01.0

0.0 0.1 0.2 0.3 0.4 0.50.0

0.2

0.4

0.6

0.8

1.0

1.2

T TF0

PP 0

Bosonsd

0.1

0.2

0.3

0.4

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

Κ kF0

PP 0

FIG. 1. (Color online) Variation of rescaled pressure with T forfour values of κ in (a) and (c), and with κ for four values of T in (b)and (d).

where b(T ,κ) is the th virial coefficient with b1(T ,

κ) = 1.We obtain the expansion for the fugacity, from inversion

of power series in Eq. (25), and write explicitly only up tothird-order terms in a parameter,

ξ = n�3

2[1 + φ(κ�)]. (41)

We then get, on writing φ ≡ φ(κ�),

z = ξ

[1− ηξ

23/2

1 + 2φ

1 + φ+ η2ξ 2

22

×{(

1 + 2φ

1 + φ

)2

− 22

33/2

1 + 3φ

1 + φ

}− η3O(ξ 3)

]. (42)

Using Eqs. (25) and (27), one has

Pβ = n

[1 − ηz

25/2

1 + 2φ

1 + φ+ η2z2

24

×{(

1 + 2φ

1 + φ

)2

− 25

35/2

1 + 3φ

1 + φ

}− η3O(z3)

]. (43)

Combination of Eqs. (42) and (43) yield, finally,

Pβ = n

[1 − ηξ

25/2

1 + 2φ

1 + φ+ η2ξ 2

23

×{(

1 + 2φ

1 + φ

)2

− 24

35/2

1 + 3φ

1 + φ

}− η3O(ξ 3)], (44)

which reduces, on putting κ = 0, to the textbook results[29,49] in powers of ξ = n�3/2 for spin- 1

2 particles.Let us substitute into Eq. (44) the expression for ξ given in

Eq. (41) and compare the resulting equation with Eq. (40). Thisleads to the inequality b2(T ,κ)/b2(T ,κ = 0) = (1 + 2φ)/(1 +φ)2 � 1 implying that b2(T ,κ) becomes less negative (pos-itive) for bosons (fermions) as soon as the SO coupling isswitched on. Since the second virial coefficient representstwo-particle interactions [29], it can be interpreted that the Weylcoupling induces repulsive (attractive) statistical interactionsin bosons (fermions). Hence the effect of the SOC is toweaken the “effective” interactions resulting in the behaviorof the pressure as depicted in Figs. 1(a)–1(d). Furthermore,the decrease (increase) in the pressure suggests that the SOCmakes the gas more (less) compressible which is supported bygeneral trends in Figs. 2(a)–2(d) wherein rescaled isothermalcompressibility KT = KT/K0

T has been plotted as a functionof κ at several temperatures with K0

T = 3/(2nε0F) as isothermal

compressibility of an ideal Fermi gas without SOC at T = 0.

053607-6

WEYL SPIN-ORBIT-COUPLING-INDUCED INTERACTIONS . . . PHYSICAL REVIEW A 88, 053607 (2013)

aFermions

2.5

2.3

2.01.81.5

0 0.005 0.01 0.015 0.020

20

40

60

80

100

120

T TF0

KTKT0

bFermions1.5

1.3

1.0

0.00 0.02 0.04 0.06 0.08 0.10

5

10

15

T TF0KTKT0

0.0

0.5

1.0

3.0

Bosons c0.2 0.3 0.4 0.5 0.6 0.7

0

2

4

6

8

10

T TF0

KTKT0

0.000.01

0.03

0.05

0.10.1

Bosons d0.30 0.35 0.40 0.45 0.50

0

0.2

0.4

0.6

0.8

1

T TF0

KTKT0103

FIG. 2. (Color online) Rescaled compressibility as a function ofT for various values of κ labeled therein.

The decrease in the Pauli pressure and consequent increasein the isothermal compressibility due to increase in Rashbacoupling strength had been found by Han and Sa de Melo [39]in the course of their study of the BCS-BEC crossover in abalanced mixture of 3D ultracold fermions. The decrease inPauli pressure has been found by Ghosh et al. [41] also.

The bosonic curves for very low κ in Fig. 2(d) are showingthe artifact of diverging KT present in an ideal Bose gas forT < T 0

C � 0.436T 0F . On the other hand, a fermionic curve has

a peak at very low temperature which becomes more and morepronounced and shifts towards lower T as κ increases [cf.Figs. 2(a) and 2(b)]. Since KT is related to the derivative ofμ, any anomalous behavior in one could be expected to haveits signature in the other. Hence we plot in Figs. 3(a)–3(d)the chemical potential μ(T ,κ) as a function of T for variousvalues of κ calculated from the implicit relation given byEq. (17). These computed values have, in fact, gone as input incalculations of all the thermodynamic quantities displayed inFigs. 1, 2, and 4. Figures 3(b) and 3(d) depict magnified graphsfor some small ranges of T to reveal the structures appearingtherein.

It can be seen in Fig. 3(d) that the nature of the curves reflectincipient Bose-Einstein condensation for very low values of κ ,just like the plots for KT displayed in Fig. 2(d), although theBose condensation is inhibited by the presence of coupling-weighted 1D character in the DOS [cf. Eq. (10)]. For fermions,the numerical values of μ(T ,κ) reveal interesting behavior inthe low-T regime. For small κ , μ(T ,κ) is increasing with adecrease in T just like in the case of an ideal 3D Fermi gaswithout SOC. However, starting at κ ≈ 0.5 k0

F, the fermionchemical potential exhibits a shoulder near T ≈ 0.12 T 0

F whichgrows upon further increase of κ and eventually forms onelocal maximum and one local minimum whose loci are beingdepicted in Fig. 3(a) by the dotted curve. The right portion(red dots) of the peak of this curve is the locus of maxima

1.01.21.5

0.6aFermions

0.7

0.85

20 0.1 0.2 0.3

0.050

0.1

0.2

0.3

0.4

0.5

T TF0

ΜF0

bFermions 1.0

2.01.5

1.2

0.00 0.05 0.10 0.15 0.200.05

0.00

0.05

0.10

T TF0

ΜF0

0.0

cBosons

0.5

1.0

3.0 1.5

0.0 0.2 0.4 0.6 0.8 1.02.0

1.5

1.0

0.5

0.0

T TF0

ΜF0 0.1 0.05 0.03 0.01 0.0

dBosons

0.0 0.1 0.2 0.3 0.4 0.51.0

0.8

0.6

0.4

0.2

0.0

0.2

T TF0

ΜF0

FIG. 3. (Color online) Variation of rescaled chemical potential asa function of T for various values of κ . Dotted curve in (a) is thelocus of finite-T extrema; starting from peak location, clockwise isfor maxima (red dots) and anticlockwise for minima (green dots).

whereas the left portion (green dots) is only a part of the locusof minima; it touches the vertical axis at μ ≈ 0.40 when κ =κc ≈ 0.63k0

F as described in Appendix C where the detailedanalysis of μ(T ,κ) versus T for successively increasing κ

starting from zero has been presented.It is worth mentioning that Ref. [51] has discussed tem-

perature variation of the chemical potential for an ideal Fermigas without SOC wherein a low-T peak has been found inless than 2D. In addition, it has been discussed in [52] thatan ideal Fermi gas confined between penetrable multilayers ormultitubes has effective reduced dimensionality and there aredimensional crossovers from 3D to 2D and then to 1D as theimpenetrability increases resulting in the appearance of a peakin μ as a function of T . Hence the presence of both extrema in acurve in our work cannot be understood solely in terms of SO-coupling-induced dimensionality reduction and its probableexplanation is given in Appendix C. In addition, it is clearfrom Eq. (C1) that fermionic μ(T ,κ) has different behaviordepending on whether the Fermi level εF(κ) ≡ μ(T = 0,κ) ofa SO-coupled system is above or below εSO.

Figures 4(a)–4(f) show temperature dependence of CV

for various values of κ . We have from Figs. 4(a) and 4(d)that (CV /NkB) → 3/2 (1/2) at high temperatures consistentwith the Dulong-Petit law for low-κ (high-κ) when thesystem is expected to have quasi-3D (quasi-1D) behavior.In addition, for T → 0,(CV /NkB) → 0 as per the third lawof thermodynamics. Figures 4(b) and 4(c) are blown upcurves in low-T regions to depict clearly the structures inthe curves. It is found from Fig. 4(c) that as T decreases, eachcurve has a dip before showing a peak at very low temperature,which is again an anomalous behavior. This kind of structurein the specific heat curve for fermions in reduced effectivedimension has been reported in Ref. [52] also. Furthermore,it can be seen that Fig. 4(e) supports SO-coupled incipient

053607-7

REENA GUPTA, G. S. SINGH, AND JURGEN BOSSE PHYSICAL REVIEW A 88, 053607 (2013)

0.0Fermions 0.5

a

1.01.53.0

0 0.1 0.2 0.30.0

0.2

0.4

0.6

0.8

1.0

1.2

T TF0

CVNk B

0.0d

0.5

1.01.53.0

Bosons

0.0 0.2 0.4 0.6 0.8 1.00.0

0.5

1.0

1.5

2.0

T TF0CVNk B

1.82.02.5

3.0

Fermions b

0 0.002 0.004 0.0060.40

0.45

0.50

0.55

0.60

0.65

T TF0

CVNk B

0.0 e

Bosons

0.010.03

0.05

0.1

0.3 0.4 0.5 0.6 0.71.65

1.70

1.75

1.80

1.85

1.90

T TF0

CVNk B

3.0

2.5

2.0

1.8

Fermionsc

0.00 0.05 0.10 0.15 0.200.52

0.54

0.56

0.58

0.60

T TF0

CVNk B

0.0

Bosons f

0.51.01.53.0

0.00 0.05 0.10 0.15 0.200.40

0.45

0.50

0.55

0.60

T TF0

CVNk B

FIG. 4. (Color online) Variation of rescaled specific heat as afunction of T for various values of κ .

Bose condensation, and crossings of the curves for various κ

in Figs. 4(b) and 4(f) are required by the Dulong-Petit lawsince the effective dimensionality changes as a consequenceof change in the coupling strength.

, , KT Fermions

, , KT Bosons

0.0 0.5 1.0 1.5 2.0 2.5 3.02.2

2.4

2.6

2.8

3.

0.45

0.5

0.54

0.58

0.63

Κ

KT

FIG. 5. (Color online) P − κ and KT − κ isotherms for harmon-ically trapped gases. Set of lower (upper) four curves are for T = 0.6(0.7). Left (right) vertical axis has values of P (KT). Symbols triangleand square (star and circle) are plots for P (KT). In addition, triangleand star (square and circle) represent bosons (fermions).

FIG. 6. (Color online) Bosonic phase diagrams (a) P − κ and(b) μ − T for values of T and κ labeling the curves. Shaded regionrepresents the Bose-condensed phase.

The isotherms P − κ and KT − κ are shown combinedin Fig. 5 for harmonically trapped noninteracting bosons andfermions; the values of the rescaled quantities P = P(κ)/P(0)and KT = KT(κ)/KT(0) are on the left and right vertical axes,respectively. We have introduced the symbol for any physicalquantity χ (κ = 0) as simply χ (0). The lower set of fourcurves starting from the lowermost, two dashed (fermions,bosons) and two solid (bosons, fermions), is for the reducedtemperature T = T/TF(0) = 0.6 and the upper set for T =0.7. The reduced coupling strength is κ = κ/kF(0). An SO-coupled trapped Bose gas shows Bose-Einstein condensation,unlike SO-coupled uniform gas, and hence temperatures havebeen chosen to correspond to the Bose gas in the normalphase for any κ . In addition, P(0) = ε4

F(0)/12h3 = nεF(0)/4and KT(0) = 3/nεF(0) with εF(0) as the Fermi energy forharmonically trapped fermions without SOC. The graphs inFig. 5 suggest that the conclusion regarding the SOC-inducedinteractions is the same here as that described earlier for uni-form gases, i.e., the increase in the coupling strength makes thestatistical interaction weaker and weaker in trapped gases, too.

Figure 6(a) shows the bosonic P − κ phase diagramwherein each curve is labeled by its value of T , the dashedcurve is the transition line for the Bose-Einstein condensation,and the shaded region represents the condensed phase. Itcan be seen that increase in the SOC strength decreases thetransition temperature which is corroborated by the μ − T

phase diagram, with μ = μ/εF(0) for various values of κ

depicted in Fig. 6(b). This kind of TC versus κ behavior isconsistent with the weakening effect of the statistical attractiveinteraction on increasing the value of κ . However, it is atvariance with the study of Hu and Liu [33] who observedthat in the thermodynamic limit, TC would not get affectedby the Rashba coupling. Our study is for the Weyl couplingand separate study for the Rashba coupling might be requiredto resolve the issue that the latter would have zero effectivestatistical interaction in trapped bosons whereas it does havefinite effective interaction in uniform Fermi gases as mentionedearlier in this section while analyzing T = 0 results of Han andSa de Melo [39] and of Ghosh et al. [41].

V. SUMMARY AND OUTLOOK

We have developed exact analytical expressions for thegrand potential in uniform as well as harmonically trapped

053607-8

WEYL SPIN-ORBIT-COUPLING-INDUCED INTERACTIONS . . . PHYSICAL REVIEW A 88, 053607 (2013)

cases in a unified form for noninteracting Bose, Fermi, andBoltzmann gases having 3D isotropic spin-orbit coupling,called the Weyl coupling. The expressions for various ther-modynamic quantities have been derived thereafter, and theircomputed values have been depicted graphically as a functionof temperature as well as the spin-orbit-coupling constant. Itis found that the Weyl coupling induces statistical interactionswhich weaken the effective interparticle interactions present inthe systems without spin-orbit coupling: Bosons have reducedeffective attraction and fermions have reduced effective repul-sion whether the gas is uniform or trapped. The conclusionhas been arrived at from graphical depiction of pressure inboth the situations; the result has been corroborated by plotsof the isothermal compressibility and analytically by the virialexpansion of the EOS in the uniform system. The variationof chemical potential and specific heat has also been analyzedand anomalous behavior revealed in the case of uniform gases.The plots of internal energy, free energy, and entropy have notbeen displayed here as they do not show any feature worthdescribing.

Our studies suggest that the physics due to syntheticspin-orbit coupling is richer than previously appreciated—not only the theoretically investigated features explainedplausibly in terms of coupling-induced enhanced DOS butsome additional features also can be understood in termsof coupling-induced modifications in interparticle statisticalinteractions. It is found that a uniform Bose gas with veryweak coupling shows incipient Bose-Einstein condensationand a Fermi system has different behaviors on varying thecoupling constant if the Fermi level is above or below theDirac point. Our work for uniform gases in conjunction withthat in Ref. [35] would constitute complete thermodynamicsof noninteracting atomic quantum gases with isotropic 3Dspin-orbit coupling. Most of the expressions derived for thethermodynamic quantities of uniform gases and all those ofharmonically trapped gases are new, and the study of completethermodynamics would be possible also for harmonicallytrapped gases with Weyl SOC in any range of physicalparameters. Moreover, we have pointed out that the resultsobtained in Refs. [33,39] for Rashba coupling are at variancewith each other if analyzed in terms of effective statisticalinteractions.

The fluctuation-dissipation theorem gives the relation [53]limq→0 S(q,T ) = nkBT KT with S(q,T ) as the temperature-dependent static structure factor. Hence the enhanced (dimin-ished) compressibility of the gas, along with any structuretherein, would get reflected in measurement of enhanced(diminished) long-wavelength fluctuation of the bunching(antibunching) effect [54,55], quantified by S(q ≈ 0,T ), inbosons (fermions). Our theoretical study of S(q,T ) andhence the pair correlation function leading to the statisticalinterparticle interaction potential will form part of a separatecommunication.

ACKNOWLEDGMENTS

R.G. acknowledges a grant of research fellowshipby the MHRD, and thanks Zeng-Qiang Yu for usefulcorrespondence.

APPENDIX A: DENSITY OF STATES FOR SO-COUPLEDUNIFORM GASES

The use of dispersion relations given in Eq. (2) gives theforms of partial DOS, ρ±(ε) ≡ ρ±(ε,κ) = (1/V )

∑k δ(ε −

ε±(k)), as

ρ+(ε) = 1

2ρ0

3D(ε)

(1 −

√εSO

ε

)2

�(ε − εSO), (A1)

and

ρ−(ε) = ρ03D(ε)

[(1 + εSO

ε

)�(εSO − ε) + 1

2

(1 +

√εSO

ε

)2

� (ε − εSO)

], (A2)

whence we have the expression for ρ(ε) as given in Eq. (10).It is found that ρ(ε) has the minimum value ρmin(ε) =

2ρ03D(εSO) at ε = εSO. The minimum arises because ρ−(ε)

decreases continuously in the range 0 � ε � ∞ but ρ+(ε) = 0for ε < εSO and increases monotonically for ε > εSO. Thevariation of rescaled DOS, ρ(ε) = ρ(ε)/ρ0

3D(ε0F), as a function

of reduced energy ε = ε/ε0F is depicted in Fig. 7(a) wherein

the dashed curve is the locus of minima in the DOS for variousvalues of κ . The plots of rescaled ρ+(ε), ρ−(ε), and ρ(ε) withrespect to ρ0

3D(ε) as a function of ε for κ = 0.5 and 1.0 areshown, respectively, in Figs. 7(c) and 7(d). Whereas the valueof ρ+(ε) goes down for increased κ and a given value of ε,the corresponding values of both ρ−(ε) and ρ(ε) go up. Inaddition, it is seen in Fig. 7(b) that ρ(ε) versus the κ curve isparabolic which gets flattened for an increased value of energy.

Locus of minimaa0.0

1.01.5

2.0 2.53.0

0 10 20 30 40 500

2

4

6

8

10

F0

ΡΡ 3D0

F0

b

1.51.00.60.30.1

2.5

0.010.0 0.2 0.4 0.6 0.8 1.00.0

0.5

1.0

1.5

2.0

Κ kF0

ΡΡ 3D0

F0

SO

Κ kF0 0.5

c

ΡΡ

Ρ

0 1 2 3 4 50.0

0.5

1.0

1.5

2.0

2.5

F0

GΡ 3D0

SO

Κ kF0 1.0

d

Ρ

Ρ

Ρ

0 1 2 3 4 50.0

0.5

1.0

1.5

2.0

2.5

F0

GΡ 3D0

FIG. 7. (Color online) Variation of rescaled DOS with ε in (a),(c), and (d), and with κ in (b). Values of κ are labeled in (a), and ε in(b). G(ε) in (c) and (d) stands for ρ+(ε), ρ−(ε) or ρ(ε). Location of ε

for minimum in ρ(ε) is marked by the arrow below which the DOSfor + helicity branch is zero.

053607-9

REENA GUPTA, G. S. SINGH, AND JURGEN BOSSE PHYSICAL REVIEW A 88, 053607 (2013)

APPENDIX B: ISOCHORIC SPECIFIC HEAT FORSO-COUPLED UNIFORM GASES

The relation CV = (∂E/∂T )V,N gives, on using Eq. (21)and the recurrence relation (19),

CV = 3

2

kBV

η�3[5Li5/2(ηz) + φ(κ�)Li3/2(ηz)]

+ V

ηβ�3[3Li3/2(ηz) + φ(κ�)Li1/2(ηz)]

1

z

(∂z

∂T

)V,N

.

(B1)

In order to find out an expression for (1/z)(∂z/∂T )V,N , werewrite Eq. (17) as

Li3/2(ηz) + φ(κ�)Li1/2(ηz) = Nη�3/2V, (B2)

and apply the operator (∂/∂T )V,N from left on both the sides.We then use in the resulting expression Eq. (19) on the left-hand side and Eq. (B2) on the right-hand side. The processgives, finally,

LHS = 1

z

(∂z

∂T

)V

[Li1/2(ηz) + φ(κ�)Li−1/2(ηz)]

− 1

Tφ(κ�)Li1/2(ηz), (B3)

and

RHS = − 3

2T[Li3/2(ηz) + φ(κ�)Li1/2(ηz)], (B4)

whence we obtain

1

z

(∂z

∂T

)V,N

= − 1

2T

[3Li3/2(ηz) + φ(κ�)Li1/2(ηz)]

[Li1/2(ηz) + φ(κ�)Li−1/2(ηz)]. (B5)

Substitution of the above expression in Eq. (B1) and thenuse of Eq. (17) finally yields expression for the specific heatas given in Eq. (28).

APPENDIX C: FERMIONIC CHEMICAL POTENTIAL

In order to analyze analytically the existence of extrema infermionic μ(T ,κ), we use the low-T approximate expression,

μ(T ,κ) = μ0

[1 − π2

12

(kBT

μ0

)2μ0 − εSO

μ0 + εSO

], (C1)

available from Ref. [35]. The location of the Dirac pointcharacterized by the energy εSO increases monotonicallywhereas the zero-temperature chemical potential μ0 ≡ μ(T =0,κ), i.e., the Fermi energy ε0

F, decreases on increasing κ .The intersection of these two curves in the ε − κ plane takesplace at the critical value κc = 2−2/3k0

F which is the solution ofμ0 = εSO, i.e., μ(T = 0,κ) = h2κ2/(2m) with the expression

for μ(T = 0,κ) given in [35]. It is clear from Eq. (C1) that for0 � κ < κc (i.e., μ0 > εSO), the curve bends down implyingthe maximum to be only at T = 0. On the other hand, the low-Tasymptote bends up for κ > κc implying that there is at leastone maximum at finite T because ultimately μ has to becomenegative at high T .

The detailed analysis of μ(T ,κ) versus T curves, generatedbased on computed values using the implicit relation for μ

given in Eq. (17), for successively increasing κ starting fromzero reveals the following. Initially, the curve has normalbehavior like an ideal Fermi gas without SOC but for κ ≈ 0.5,a shoulder begins to develop at T ≈ 0.12T 0

F which gets moreand more pronounced until κ ≈ 0.5985 but μ has its maximumat T = 0 for all these values of κ . On increasing κ , theso far developing “hanging shoulder” turns into an “athleticshoulder” showing a local minimum at Tmin > 0 and a localmaximum at Tmax > Tmin apart from the global maximum atT = 0. We note that we get μ(Tmax,κ)/μ(0,κ) < 1 on the onsetof the max-min-max structure and this continues until theinequality gets reversed to μ(Tmax,κ)/μ(0,κ) > 1 for κ ≈ 0.61without any change in order of the max-min-max structure.The finite-T minimum (maximum) wanders to lower (higher)T as κ is increased further. Finally, at κ = κc, the finite-Tminimum gets merged with the zero-T maximum leaving alocal minimum at T = 0 and a global maximum at finite T .This structure continues for higher values of κ but the locationof the global maximum goes on shifting towards the verticalaxis being depicted in Fig. 3(a) as the right portion (red dots)of the dotted curve.

What is the physical implication of ∂μ

∂T> 0 for κ > κc?

μ(T ,κ) is the energy necessary to add one fermion to thesystem. For an ideal gas, this energy is the larger; the lower is T .It is the largest at T = 0, where it takes ε0

F to add an additionalfermion to the system. It takes so much energy even in thenoninteracting gas, because there is the Pauli principle actinglike an effective repulsion between particles. This statistics-induced repulsion is softened by SOC. To add a fermion atT = 0 now will be the easier; the larger is κ . This statementalso holds for any fixed T : μ(T ,κ1) > μ(T ,κ2), if κ2 > κ1.On the contrary, for fixed κ , a corresponding statementdoes hold only for κ < κc: μ(T1,κ) > μ(T2,κ), if T2 > T1

and κ < κc.Obviously, the finite-T maximum and its precursors (shoul-

der, min-max) result from a competition between the effectsof the statistical repulsion, which continuously increases withdecreasing T , and the SOC, the effect of which is amplified bya factor 1/T and hence enormously at low T . In addition, forlarge κ (fixed T ), the distinction between bosons and fermionsgets lost similar to what we find for large T in ideal quantumgases without SOC.

[1] J. Dalibard, F. Gerbier, G. Juzeliunas, and P. Ohberg, Rev. Mod.Phys. 83, 1523 (2011).

[2] Z. F. Xu and L. You, Phys. Rev. A 85, 043605 (2012).[3] E. I. Rashba, Fiz. Tverd. Tela 2, 1224 (1960) [Sov. Phys.-Solid

State 2, 1109 (1960)]; G. Dresselhaus, Phys. Rev. 100, 580(1955).

[4] Xiang-Fa Zhou, Guang-Can Guo, Wei Zhang, and Wei Yi, Phys.Rev. A 87, 063606(R) (2013).

[5] Y.-J. Lin, K. Jimenez-Garcıa, and I. B. Spielman, Nature(London) 471, 83 (2011).

[6] J.-Y. Zhang, S.-C. Ji, Z. Chen, L. Zhang, Z.-D. Du,B. Yan, G.-S. Pan, B. Zhao, Y.-J. Deng, H. Zhai,

053607-10

WEYL SPIN-ORBIT-COUPLING-INDUCED INTERACTIONS . . . PHYSICAL REVIEW A 88, 053607 (2013)

S. Chen, and J.-W. Pan, Phys. Rev. Lett. 109, 115301(2012).

[7] P. Wang, Z.-Q. Yu, Z. Fu, J. Miao, L. Huang, S. Chai, H. Zhai,and J. Zhang, Phys. Rev. Lett. 109, 095301 (2012); L. W. Cheuk,A. T. Sommer, Z. Hadzibabic, T. Yefsah, W. S. Bakr, and M. W.Zwierlein, ibid. 109, 095302 (2012).

[8] Zhengkun Fu, Lianghui Huang, Zengming Meng, PengjunWang, Xia-Ji Liu, Han Pu, Hui Hu, and Jing Zhang, Phys. Rev.A 87, 053619 (2013).

[9] B. M. Anderson, G. Juzeliunas, V. M. Galitski, and I. B.Spielman, Phys. Rev. Lett. 108, 235301 (2012).

[10] Yi Li, Xiangfa Zhou, and Congjun Wu, Phys. Rev. B 85, 125122(2012).

[11] X. Wan, A. M. Turner, A. Vishwanath, and S. Y. Savrasov, Phys.Rev. B 83, 205101 (2011); A. A. Burkov and Leon Balents, Phys.Rev. Lett. 107, 127205 (2011).

[12] M. Maggiore, A Modern Introduction to Quantum Field Theory(Oxford University Press, Oxford, 2005).

[13] Xiangfa Zhou, Yi Li, Zi Cai, and Congjun Wu, J. Phys. B 46,134001 (2013).

[14] Hui Zhai, Int. J. Mod. Phys. B 26, 1230001 (2012); VictorGalitski and Ian B. Spielman, Nature 494, 49 (2013).

[15] Subhasis Sinha, Rejish Nath, and Luis Santos, Phys. Rev. Lett.107, 270401 (2011); Hui Hu, B. Ramachandhran, Han Pu, andXia-Ji Liu, ibid. 108, 010402 (2012).

[16] Chunji Wang, Chao Gao, Chao-Ming Jian, and Hui Zhai, Phys.Rev. Lett. 105, 160403 (2010); Tin-Lun Ho and Shizhong Zhang,ibid. 107, 150403 (2011).

[17] R. P. Feynman, Statistical Mechanics, A Set of Lectures.(Addison-Wesley, Reading, 1972).

[18] Conjun Wu, Mod. Phys. Lett. B 23, 1 (2009); Cong-Jun Wu, IanMondragon-Shem, and Xiang-Fa Zhou, Chin. Phys. Lett. 28,097102 (2011).

[19] Chunlei Qu, Chris Hamner, Ming Gong, Chuanwei Zhang, andPeter Engels, Phys. Rev. A 88, 021604(R) (2013).

[20] M. Gong, S. Tewari, and C. Zhang, Phys. Rev. Lett. 107, 195303(2011).

[21] Yi Li and Congjun Wu, Phys. Rev. B 85, 205126 (2012); Xiong-Jun Liu, K. T. Law, and T. K. Ng, arXiv:1304.0291.

[22] J. P. Vyasanakere and V. B. Shenoy, Phys. Rev. B 83, 094515(2011).

[23] H. Hu, L. Jiang, X.-J. Liu, and H. Pu, Phys. Rev. Lett. 107,195304 (2011).

[24] Zeng-Qiang Yu and Hui Zhai, Phys. Rev. Lett. 107, 195305(2011).

[25] M. Sato, Y. Takahashi, and S. Fujimoto, Phys. Rev. Lett. 103,020401 (2009); Ming Gong, Gang Chen, Suotang Jia, andChuanwei Zhang, ibid. 109, 105302 (2012); X.-J. Liu and P. D.Drummond, Phys. Rev. A 86, 035602 (2012); R. Wei and E. J.Mueller, ibid. 86, 063604 (2012); Chunlei Qu, Zhen Zheng,Ming Gong, Yong Xu, Li Mao, Xubo Zou, Guangcan Guo, andChuanwei Zhang, arXiv:1307.1207v2.

[26] Fan Wu, Guang-Can Guo, Wei Zhang, and Wei Yi, Phys. Rev.Lett. 110, 110401 (2013); Zhen Zheng, Ming Gong, XuboZou, Chuanwei Zhang, and Guangcan Guo, Phys. Rev. A87, 031602(R) (2013); M. Iskin and A. L. Subasi, ibid. 87,063627 (2013); Xia-Ji Liu and Hui Hu, ibid. 88, 023622 (2013);Fan Wu, Guang-Can Guo, Wei Zhang, and Wei Yi, ibid. 88,043614 (2013); Zhen Zheng, Ming Gong, Yichao Zhang, XuboZou, Chuanwei Zhang, and Guangcan Guo, arXiv:1212.6826;

Chunlei Qu, Zhen Zheng, Ming Gong, Yong Xu, Li Mao, XuboZou, Guangcan Guo, and Chuanwei Zhang, arXiv:1307.1207;Wei Zhang and Wei Yi, arXiv:1307.2439.

[27] G.-B. Zhu, Q. Sun, Y.-Y. Zhang, K. S. Chan, W.-M. Liu, andA.-C. Ji, Phys. Rev. A 88, 023608 (2013).

[28] Immanuel Bloch, Jean Dalibard, and Wilhelm Zwerger, Rev.Mod. Phys. 80, 885 (2008); S. Giorgini, L. P. Pitaevskii, andS. Stringari, ibid. 80, 1215 (2008); M. Inguscio, W. Ketterle,and C. Salomon (eds.), Proceedings of the International Schoolof Physics “Enrico Fermi,” Course CLXIV, Varenna, 20–30June 2006 (IOS Press, Amsterdam, 2008); Immanuel Bloch,Jean Dalibard, and Sylvain Nascimbene, Nat. Phys. 8, 267(2012).

[29] L. D. Landau and E. M. Lifshitz, Statistical Mechanics, Part I,3rd ed. (Pergamon Press, Oxford, 1993).

[30] There is at least one exception to it. Spin-polarized 1H atomsare sometimes referred to as “spin- 1

2 bosons,” see, for example,A. J. Legget, Quantum Liquids: Bose Condensation and CooperPairing in Condensed-Matter Systems (Oxford University Press,New York, 2006), p. 8.

[31] W. Pauli, Phys. Rev. 58, 716 (1940).[32] Xiaoling Cui and Qi Zhou, Phys. Rev. A 87, 031604(R)

(2013).[33] H. Hu and Xia-Ji Liu, Phys. Rev. A 85, 013619 (2012).[34] Tomoki Ozawa and Gordon Baym, Phys. Rev. Lett. 109, 025301

(2012).[35] Li He and Zeng-Qiang Yu, Phys. Rev. A 84, 025601

(2011).[36] J. P. Vyasanakere, S. Zhang, and V. B. Shenoy, Phys. Rev. B 84,

014512 (2011).[37] Qi Zhou and Xiaoling Cui, Phys. Rev. Lett. 110, 140407

(2013).[38] Wei Zheng, Zeng-Qiang Yu, Xiaoling Cui, and Hui Zhai, J. Phys.

B 46, 134007 (2013).[39] Li Han and C. A. R. Sa de Melo, Phys. Rev. A 85, 011606(R)

(2012).[40] Zhenhua Yu, Phys. Rev. A 85, 042711 (2012).[41] Sudeep Kumar Ghosh, Jayantha P. Vyasanakere, and Vijay B.

Shenoy, Phys. Rev. A 84, 053629 (2011).[42] Brandon M. Anderson and Charles W. Clark, J. Phys. B 46,

134003 (2013); although the paper deals with the harmonicallytrapped gas with the SOC, it is due to isotropic trapping that thedescribed simplifying aspects would apply straightway in thecase of SO-coupled uniform gases, too.

[43] Vijay B. Shenoy, Jayantha P. Vyasanakere, and S. K. Ghosh,Current Science 103, 1 (2012).

[44] Vanderlei Bagnato, David E. Pritchard and Daniel Kleppner,Phys. Rev. A 35, 4354 (1987).

[45] I. S. gradeshteyn and I. M. Ryzhik, Table of Integrals, Seriesand Products, 7th ed. (Academic Press, New York, 2007).

[46] Luca Salasnich, J. Math. Phys. 41, 8016 (2000).[47] A. P. Prudnikov, Yu. A. Brychkov, and O. I. Marichev, Integrals

and Series, Volume 3: More Special Functions (Gordon andBreach, New York, 1990).

[48] M. H. Lee, Acta Physica Polonica B 40, 1279 (2009).[49] R. K. Pathria and Paul D. Beale, Statistical Mechanics, 3rd ed.

(Butterworth-Heinemann, Oxford, 2011).[50] See, for example, Vıctor Romero-Rochın, Phys. Rev. Lett.

94, 130601 (2005); Vıctor Romero-Rochın and Vanderlei S.Bagnato, Brazilian J. Phys. 35, 607 (2005).

053607-11

REENA GUPTA, G. S. SINGH, AND JURGEN BOSSE PHYSICAL REVIEW A 88, 053607 (2013)

[51] M. Grether, M. de Llano, and M. A. Solıs, Eur. Phys. J. D 25,287 (2003).

[52] P. Salas and M. A. Solıs, arXiv:1307.2649.[53] L. Pitaevskii and S. Stringari, Bose–Einstein Condensation

(Clarendon Press, Oxford, 2003).

[54] J. Bosse, K. N. Pathak, and G. S. Singh, Phys. Rev. E 84, 042101(2011).

[55] A. Blumkin, S. Rinott, R. Schley, A. Berkovitz,I. Shammass, and J. Steinhauer, Phys. Rev. Lett. 110, 265301(2013).

053607-12