This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Effective electromagnetic actions for Lorentz violating theories exhibiting the axial anomaly

Andrés Gómez [email protected] Facultad de Ciencias, Universidad Nacional Autónoma de México,
04510 México, Distrito Federal, México
   A. Martín-Ruiz [email protected]    Luis F. Urrutia [email protected] Instituto de Ciencias Nucleares, Universidad Nacional Autónoma de México,
04510 México, Distrito Federal, México
Abstract

The CPT odd contribution to the effective electromagnetic action deriving from the vacuum polarization tensor in a large class of fermionic systems exhibiting Lorentz invariance violation (LIV) is calculated using thermal field theory methods, focusing upon corrections depending on the chemical potential. The systems considered exhibit the axial anomaly and their effective actions are described by axion electrodynamics whereby all the LIV parameters enter in the coupling Θ(x)\Theta(x) to the unmodified Pontryagin density. A preliminary application to type-I tilted Weyl semimetals is briefly presented.

Lorentz violation, Vacuum polarization, Thermal field theory, Axial anomaly

I Introduction

The study of quantum corrections in the QED fermionic sector of the Standard Model Extension (SME) Kostelecky0 ; Kostelecky1 due to the electromagnetic interaction was sparkled in Ref. Kostelecky1 which posed the problem of obtaining corrections to the Chern-Simons interaction 12ϵμναβΔkμAνFαβ\frac{1}{2}\epsilon^{\mu\nu\alpha\beta}{\Delta k}_{\mu}A_{\nu}F_{\alpha\beta} as arising from the one loop vacuum polarization tensor of fermions with the additional coupling Ψ¯γ5γμbμΨ{\bar{\Psi}}\gamma^{5}\gamma^{\mu}b_{\mu}\Psi. Since then, a large number of authors have carried the calculation of the CPT odd contribution to the vacuum polarization obtaining a result that can be presented as Δkμ=ζ(e2/π2)bμ\Delta k^{\mu}=\zeta\,(e^{2}/\pi^{2})\,b^{\mu}, where ζ\zeta is always a finite coefficient adopting different values, including ζ=0\zeta=0, according to the regularization prescription chosen to deal with the superficially linear divergence of the vacuum polarization tensor Πμν\Pi^{\mu\nu} CG ; JK ; PV ; CHUNG1 ; CHUNG2 ; CHEN ; BONNEAU ; CHAICHIAN ; PV1 ; BATTISTEL ; ANDRIANOV ; BALT1 ; BALT2 ; ALFARO . A survey of the principal approaches regarding this issue can be found in Refs. PV1 ; ALFARO ; JACKIW2 ; BALT3 . The search of a physical criterion to select one specific value of ζ\zeta has been unsuccessful and the current understanding is that ζ\zeta would be fixed either by an experimental condition or by a more fundamental theory. In order to briefly summarize some of the previous results we assume that bμb_{\mu} is timelike. The case ζ=0\zeta=0 is obtained by demanding gauge invariance of the effective Lagrangian density CG , and has the drawback of eliminating the Chern-Simons contribution from the outset since this term changes by a total derivative under gauge transformations. Using the full non-perturbative expression in bμb_{\mu} for the fermion propagator and different regulators consistent with symmetric integration in four dimensions, the authors of Refs. JK ; PV ; CHUNG1 ; CHUNG2 find |ζ|=3/16|\zeta|={3}/{16}. In an attempt to elucidate a further significance of the non-perturbative approach, which demands the use of a unique regulator for all the contributions of Πμν\Pi^{\mu\nu}, the CPT even contribution to Πμν\Pi^{\mu\nu} (second order in bμb_{\mu}) was calculated in the m=0m=0 case, finding that gauge invariance is violated BALT1 . The use of Pauli-Villars regularization was subsequently introduced in the non-perturbative calculation of the massive case, yielding gauge invariance to second order but two possibilities for the CPT odd contribution: either the expected value ζ=0\zeta=0 or |ζ|=3/8|\zeta|=3/8 BALT2 . Again, there is no unique way of fixing the ζ\zeta contribution by demanding gauge invariance of Πμν\Pi^{\mu\nu}. An alternative regularization prescription based upon the maximal residual symmetry group of the vector bμb_{\mu}, proposed in Ref. ANDRIANOV , provides a definition on how to perform the loop integration d4kd^{4}k in Πμν\Pi^{\mu\nu}. For example, in the timelike case, the maximal residual symmetry group is SO(3)SO(3) which calls for a spherically symmetric integration in the tri-momentum 𝐤{\mathbf{k}} instead of on the four-momentum kμk^{\mu}. This prescription yields |ζ|=1/4|\zeta|=1/4.

In this work we extend previous calculations of effective electromagnetic action induced by radiative correction in the SME by including additional terms of the fermionic sector in the minimal QED extension of the SME TABLE . Our starting point is the action

S=d4xΨ¯(ΓμiμMeΓμAμ)Ψ,\displaystyle S=\int d^{4}x\,\bar{\Psi}\left(\Gamma^{\mu}i\partial_{\mu}-M-e\Gamma^{\mu}A_{\mu}\right)\Psi, (1)

which is coupled to the electromagnetic field Aμ=(A0,Ai)=(A0,𝑨)A^{\mu}=(A^{0},A^{i})=(A^{0},{\mbox{\boldmath$A$}}), with the notation 𝑨=(Ax,Ay,Az){\mbox{\boldmath$A$}}=(A_{x},\,A_{y},\,A_{z}) in terms of the Cartesian components. The metric is ημν=diag(+,,,)\eta_{\mu\nu}=\mbox{diag}(+,-,-,-) and we use =c=1\hbar=c=1 units henceforth. We set m=m5=Hμν=eλ=fλ=gλκν=0m=m_{5}=H_{\mu\nu}=e_{\lambda}=f_{\lambda}=g_{\lambda\kappa\nu}=0, in the notation of Table XVI of Ref. TABLE . Thus we restrict ourselves to

Γμ=cμγνν+dμγ5νγν,M=aμγμ+bμγ5γμ,\Gamma^{\mu}=c^{\mu}{}_{\nu}\gamma^{\nu}+d^{\mu}{}_{\nu}\gamma^{5}\gamma^{\nu},\qquad M=a_{\mu}\gamma^{\mu}+b_{\mu}\gamma^{5}\gamma^{\mu}, (2)

with γμ\gamma^{\mu} being the standard gamma matrices. The term cμνc^{\mu}{}_{\nu} already includes the δνμ\delta^{\mu}_{\nu} contribution corresponding to the free Dirac action. The terms Ψ¯ΓμiμΨ{\bar{\Psi}}\,\Gamma^{\mu}i\partial_{\mu}\Psi are CPT even, while those in Ψ¯MΨ{\bar{\Psi}}M\Psi are CPT odd. Still, since γ5\gamma^{5} is PT odd, each of them separately contains a mixture of PT even and odd terms.

Let us emphasize that the methods and techniques employed in the calculation of these extended effective actions in high energy physics, besides being valuable in their own, can be of relevance in some areas of condensed matter physics. Indeed, the identification of fermionic quasiparticles of Dirac and/or Weyl type in the linearized approximation of Hamiltonians in topological phases of matter provides the opportunity of studying them under the perspective of the SME. Such approach has been particularly fruitful in the case of Weyl semimetals (WSMs) whose electronic Hamiltonians naturally include some of the LIV terms considered in the fermion sector of the SME. Nevertheless, in this case, the LIV parameters need not be highly suppressed, since they are determined by the electronic structure of the material and are subjected to experimental determination GRUSHIN ; Burkov ; Goswami ; TI ; REVLANDS ; Kostelecky3 . Our choice of the LIV parameters in Eq. (2) are those relevant for the description of a general WSM. The simplest example arises in the Hamiltonian of a Weyl semimetal with no tilting and with an isotropic Fermi velocity, which can be embedded in the fermionic action

S=d4xΨ¯(x)(iγμμb~μγ5γμ)Ψ(x),\displaystyle S=\int d^{4}x\,{\bar{\Psi}}(x)\left(i\gamma^{\mu}\partial_{\mu}-{{\tilde{b}}_{\mu}\gamma^{5}\gamma^{\mu}}\right)\Psi(x), (3)

where b~μ=(b~0,𝒃~){\tilde{b}}^{\mu}=({\tilde{b}}^{0},{\mbox{\boldmath${\tilde{b}}$}}). Starting from the action (3) coupled to an electromagnetic field, the chiral rotation method of Ref. Burkov , which provides an alternative quite different from the standard vacuum polarization calculation, yields

Seff=e232π2d4xΘ(x)ϵαβμνFαβFμν,Θ(x)=2b~μxμ,\displaystyle S_{{\rm{eff}}}=\frac{e^{2}}{32\pi^{2}}\int d^{4}x\,\Theta(x)\,\epsilon^{\alpha\beta\mu\nu}F_{\alpha\beta}F_{\mu\nu},\quad\Theta(x)=2{\tilde{b}}_{\mu}x^{\mu}, (4)

which determines |ζ|=1/4|\zeta|=1/4. Another outstanding example relevant in condensed matter is the case of the superfluid 3He-A, where the value ζ=1/2\zeta=1/2 is reported VOLOVIK .

The action (4) captures the electromagnetic response of WSMs. While the axion coupling Θ(x)\Theta(x) arises from the nontrivial topology of the band structure of the material, the appearance of the abelian Pontryagin density (APD) ϵαβμνFαβFμν\epsilon^{\alpha\beta\mu\nu}F_{\alpha\beta}F_{\mu\nu} is related to the axial anomaly

μJμ5=e216π2ϵαβμνFαβFμν.\displaystyle\partial^{\mu}J_{\mu}^{5}=-\frac{e^{2}}{16\pi^{2}}\,\epsilon^{\alpha\beta\mu\nu}F_{\alpha\beta}F_{\mu\nu}. (5)

From Eq. (4) we obtain the effective sources FRANZ

ρ=δSeffδA0=e22π2𝒃~𝑩,𝑱=δSeffδ𝑨=e22π2(𝒃~×𝑬b~0𝑩),\rho=\frac{\delta S_{{\rm eff}}}{\delta A_{0}}=\frac{e^{2}}{2\pi^{2}}{\mbox{\boldmath$\tilde{b}$}}\cdot{\mbox{\boldmath$B$}},\qquad{\mbox{\boldmath$J$}}=\frac{\delta S_{{\rm eff}}}{\delta{\mbox{\boldmath$A$}}}=\frac{e^{2}}{2\pi^{2}}({\mbox{\boldmath$\tilde{b}$}}\times{\mbox{\boldmath$E$}}-{\tilde{b}}_{0}{\mbox{\boldmath$B$}}), (6)

where 𝑬E and 𝑩B are the electromagnetic fields. In our index notation the above current is Ji=δSeff/δAiJ_{i}=\delta S_{\rm eff}/\delta A^{i}. In particular, for b~0=0{\tilde{b}}_{0}=0, this yields Jx=σxyEyJ_{x}=\sigma_{xy}E_{y} with the transverse anomalous Hall conductivity σxy=e22π2b~z\sigma_{xy}=-\frac{e^{2}}{2\pi^{2}}{\tilde{b}}_{z}, a distinguishing feature of Weyl semimetals BURKOVBALENTS . As it will become clear in Section III.1, the knowledge of this quantity allows to unambiguously fix the otherwise arbitrary coefficient ζ\zeta in this case. This additional information is a general consequence of the underlying microscopic theory describing the material, as opposed to the lack of a more fundamental theory containing the SME.

Recently, the appearance of the APD in effective actions arising from (1) has promoted the use of anomaly calculations to obtain them. For example, in the path integral approach, SeffS_{\rm eff} was obtained by introducing the electromagnetic coupling in Eq. (3) and subsequently eliminating the fermionic term proportional to b~μ{\tilde{b}}_{\mu} through a chiral rotation. Nevertheless, this produces an electromagnetic contribution to the action arising from the nonzero Jacobian of the chiral rotation which is proportional to the Pontryagin density Burkov . Following this idea, the Fujikawa prescription to obtain the chiral anomalies Fujikawa ; Bertlement has also been used to calculate the effective electromagnetic action of different materials in Refs. Goswami ; TI ; REVLANDS as well as to give an alternative explanation of the indeterminacy of the coefficient ζ\zeta related to some freedom in the definition of the fermionic axial vector current CHUNG3 .

Nevertheless, as pointed out in Ref. LUAGA , the anomaly does not directly yield the effective action. Also, the method of eliminating the modifying fermionic contributions via a chiral rotation cannot be easily extended to deal with the more complicated configurations envisaged in the action (1).

The reasons indicated above suggest the convenience of applying and extending the quantum field theory methods developed in high energy physics to obtain the required effective electromagnetic actions corresponding to fermionic systems described by the action (1) which are of interest in condensed matter physics. In particular this procedure should clarify how the LIV corrections enter in the effective action, while the chiral anomaly remains insensitive to them LUAGA ; FIDEL ; SALVIO ; Scarpelli . Motivated by the inclusion of temperature effects in the odd contribution of the vacuum polarization tensor arising from the bμb_{\mu} coupling in the action (3), reported in Refs. EBERT ; TEMP1 ; TEMP2 ; TEMP3 ; TEMP4 ; TEMP5 , we provide the first steps to incorporate thermal field theory in the case of the more general LIV couplings defined in Eqs. (2). In this way, we incorporate non-zero chemical potential effects, but still remain in the zero temperature limit. We follow the conventions of Ref. PESKIN .

II The effective action

Integrating the fermions in Eq. (1) defines the effective action

exp(iSeff)=det[Ψ¯(x)(ΓμiμMeΓμAμ)Ψ(x)],\displaystyle\exp(iS_{\rm eff})=\det\Big{[}{\bar{\Psi}}(x)\,\Big{(}\Gamma^{\mu}i\partial_{\mu}-M-e\Gamma^{\mu}A_{\mu}\Big{)}\,\Psi(x)\Big{]}, (7)

which we write as

Seff(2)(A)=12d4p(2π)4Aμ(p)Πμν(p)Aν(p),\displaystyle S_{\rm eff}^{(2)}(A)=\frac{1}{2}\int\frac{d^{4}p}{\left(2\pi\right)^{4}}\,A_{\mu}(-p)\,\Pi^{\mu\nu}(p)\,A_{\nu}(p), (8)

to second order in the electromagnetic potential. This introduces the vacuum polarization tensor Πμν(p)\Pi^{\mu\nu}(p)

iΠμν(p)=e2d4k(2π)4tr[S(kp)ΓμS(k)Γν],\displaystyle i\Pi^{\mu\nu}(p)=e^{2}\int\frac{d^{4}k}{\left(2\pi\right)^{4}}\,\mbox{tr}\left[S(k-p)\Gamma^{\mu}S(k)\Gamma^{\nu}\right], (9)

where S(k)=i/(ΓμkμM)S(k)=i/(\Gamma^{\mu}k_{\mu}-M) is the exact fermion propagator in momentum space including LIV modifications. Since Seff(2)S_{\rm eff}^{(2)} is real we must have Πμν(p)=Πνμ(p)\Pi^{*}_{\mu\nu}(p)=\Pi_{\nu\mu}(p). As we will show, the new effective action (8) will keep the form of Eq. (4) with all the modifications entering through a new vector λ{\cal B}_{\lambda} to be determined. The link between both expressions is accomplished with the identifications

Πμν(p)=ie22π2λpκϵμνλκ,Θ(x)=2λxλ.\displaystyle\Pi^{\mu\nu}(p)=-i\frac{e^{2}}{2\pi^{2}}\,{\cal B}_{\lambda}p_{\kappa}\epsilon^{\mu\nu\lambda\kappa},\quad\Theta(x)=2{\cal B}_{\lambda}x^{\lambda}. (10)

In the following we calculate the CPT odd contribution of the effective action (8). In the massless case (m=0m=0), ΓμkμM\Gamma^{\mu}k_{\mu}-M is linear in γμ\gamma^{\mu} and γ5γμ\gamma^{5}\gamma^{\mu}. The appearance of the matrix γ5\gamma^{5} suggests the convenience of using left and right chiral projectors in order to evaluate γ5=±1\gamma^{5}=\pm 1 BALT1 ; SALVIO ; TEMP1 . Therefore, we can work in the chiral basis, where the decomposition of the Dirac spinor into the right and left Weyl spinors is manifest. These latter are eigenspinors of γ5\gamma^{5} with the eigenvalues ±1\pm 1. We now define the projection operators

PR=1+γ52,PL=1γ52,γ52=1,\displaystyle P_{R}=\frac{1+\gamma^{5}}{2},\qquad P_{L}=\frac{1-\gamma^{5}}{2},\qquad\gamma_{5}^{2}=1, (11)

which project onto right- and left-handed spinors, respectively. Note that γμPL=PRγμ\gamma^{\mu}P_{L}=P_{R}\gamma^{\mu}. The projectors (11) allow us to define the matrices ΓRμ\Gamma_{R}^{\mu} such that

ΓμPR=(cμνdμ)νγνPRΓRμPR,\displaystyle\Gamma^{\mu}P_{R}=\left(c^{\mu}{}_{\nu}-d^{\mu}{}_{\nu}\right)\gamma^{\nu}P_{R}\equiv\Gamma_{R}^{\mu}P_{R}, (12)

which identifies ΓRμ=rμγνν\Gamma^{\mu}_{R}=r^{\mu}{}_{\nu}\gamma^{\nu} with rμ=νcμνdμνr^{\mu}{}_{\nu}=c^{\mu}{}_{\nu}-d^{\mu}{}_{\nu}. In an analogous way we define the left-handed part of the matrices Γμ\Gamma^{\mu} as

ΓμPL=ΓLμPL,ΓLμ=lμγνν,lμ=νcμ+νdμ.ν\displaystyle\Gamma^{\mu}P_{L}=\Gamma_{L}^{\mu}P_{L},\qquad\Gamma_{L}^{\mu}=l^{\mu}{}_{\nu}\gamma^{\nu},\qquad l^{\mu}{}_{\nu}=c^{\mu}{}_{\nu}+d^{\mu}{}_{\nu}\,. (13)

Following the same idea, we can also split the fermion propagator S(k)S(k) into its right- and left-handed parts, i.e.

iΓμkμMγαPR=PRSR(k)γα,SR(k)=i(kμrμνaν+bν)γν,\displaystyle\frac{i}{\Gamma^{\mu}k_{\mu}-M}\gamma^{\alpha}P_{R}=P_{R}S_{R}(k)\gamma^{\alpha},\quad S_{R}(k)=\frac{i}{\big{(}k_{\mu}r^{\mu}{}_{\nu}-a_{\nu}+b_{\nu}\big{)}\gamma^{\nu}},
iΓμkμMγαPL=PLSL(k)γα,SL(k)=i(kμlμνaνbν)γν.\displaystyle\frac{i}{\Gamma^{\mu}k_{\mu}-M}\gamma^{\alpha}P_{L}=P_{L}S_{L}(k)\gamma^{\alpha},\quad S_{L}(k)=\frac{i}{\big{(}k_{\mu}l^{\mu}{}_{\nu}-a_{\nu}-b_{\nu}\big{)}\gamma^{\nu}}. (14)

Note that the propagators SLS_{L} and SRS_{R}, having the generic form i/(Zνγν)i/(Z_{\nu}\gamma^{\nu}), can be readily rationalized as i(Zνγν)/Z2i(Z_{\nu}\gamma^{\nu})/Z^{2}.

To proceed forward with the calculation we now split the combination Tμν(k,p)=S(kp)ΓμS(k)ΓνT^{\mu\nu}(k,p)=S(k-p)\Gamma^{\mu}S(k)\Gamma^{\nu} under the trace in Πμν(p)\Pi^{\mu\nu}(p) into its left- and right-handed parts, i.e.

TL(R)μν(k,p)\displaystyle T_{L(R)}^{\mu\nu}(k,p) =S(kp)ΓμS(k)ΓνPL(R),\displaystyle=S(k-p)\Gamma^{\mu}S(k)\Gamma^{\nu}P_{L(R)}, (15)

which implies that the vacuum polarization can be written as the sum Πμν(p)=ΠLμν(p)+ΠRμν(p)\Pi^{\mu\nu}(p)=\Pi_{L}^{\mu\nu}(p)+\Pi_{R}^{\mu\nu}(p), where

iΠL(R)μν(p)\displaystyle i\Pi_{L(R)}^{\mu\nu}(p) =e2d4k(2π)4tr[TL(R)μν(k,p)]\displaystyle=e^{2}\int\frac{d^{4}k}{\left(2\pi\right)^{4}}\,\mbox{tr}\left[T_{L(R)}^{\mu\nu}(k,p)\right] (16)

is the vacuum polarization for a left-(right-)handed massless fermion. We now concentrate in the calculation of ΠL(R)μν(p)\Pi_{L(R)}^{\mu\nu}(p). Using Eqs. (12)-(14) and the cyclic property of the trace, the left-handed part ΠLμν\Pi_{L}^{\mu\nu} can be written as

iΠLμν(p)\displaystyle i\Pi_{L}^{\mu\nu}(p) =e2lμlνβαd4k(2π)4tr[SL(kp)γβSL(k)γαPL].\displaystyle=e^{2}l^{\mu}{}_{\beta}l^{\nu}{}_{\alpha}\int\frac{d^{4}k}{\left(2\pi\right)^{4}}\,\mbox{tr}\left[S_{L}(k-p)\gamma^{\beta}S_{L}(k)\gamma^{\alpha}P_{L}\right]. (17)

A similar procedure yields the right-handed part ΠRμν(p)\Pi^{\mu\nu}_{R}(p) by means of the replacements LRL\to R, lμνrμνl^{\mu}{}_{\nu}\to r^{\mu}{}_{\nu}, bμbμb_{\mu}\to-b_{\mu}, PLPRP_{L}\to P_{R} in Eq. (17). In the following we restrict ourselves to the axial contributions ΠA,Lμν\Pi^{\mu\nu}_{A,L} and ΠA,Rμν\Pi^{\mu\nu}_{A,R} of the left- and right-handed terms, respectively, which are obtained by isolating the terms PR+γ5/2P_{R}\to+\gamma^{5}/2 and PLγ5/2P_{L}\to-\gamma^{5}/2 in the corresponding expressions for the vacuum polarization. Both expressions can be summarized in the general form

iΠA,χμν(p)\displaystyle i\Pi_{A,\,\chi}^{\mu\nu}(p) =χ2e2mμmνβαd4k(2π)4tr[Sχ(kp)γβSχ(k)γαγ5],\displaystyle=\frac{{\chi}}{2}e^{2}m^{\mu}{}_{\beta}m^{\nu}{}_{\alpha}\int\frac{d^{4}k}{\left(2\pi\right)^{4}}\mbox{tr}\left[S_{{\chi}}(k-p)\gamma^{\beta}S_{{\chi}}(k)\gamma^{\alpha}\gamma^{5}\right], (18)

with the following assignments

R\displaystyle R :\displaystyle: χ=+1mμ=νrμ,νCρ=aρbρ,\displaystyle{\chi=+1}\qquad m^{\mu}{}_{\nu}=r^{\mu}{}_{\nu},\qquad C_{\rho}=a_{\rho}-b_{\rho},
L\displaystyle L :\displaystyle: χ=1mμ=νlμ,νCρ=aρ+bρ.\displaystyle{\chi=-1}\qquad m^{\mu}{}_{\nu}=l^{\mu}{}_{\nu},\qquad C_{\rho}=a_{\rho}+b_{\rho}. (19)

Clearly, the full axial contribution to the vacuum polarization is the sum of the LL and RR parts, i.e. ΠAμν=ΠA,Lμν+ΠA,Rμν\Pi_{A}^{\mu\nu}=\Pi_{A,L}^{\mu\nu}+\Pi_{A,R}^{\mu\nu}. For the sake of simplicity, we do not introduce an additional subindex χ\chi either in the matrix mαβm^{\alpha}{}_{\beta} or in the vector CρC_{\rho}, which are to be restored at the end of the calculations according to Eq. (19). The calculation in Eq. (18) proceeds as follows: to simplify the notation we introduce the primed vectors qν=qμmμνq_{\nu}^{\prime}=q_{\mu}m^{\mu}{}_{\nu} maintaining the original measure d4kd^{4}k. Also we rationalize the denominators in the propagators and take the trace tr(γργβγσγαγ5)=4iϵρβσα\mbox{tr}(\gamma^{\rho}\gamma^{\beta}\gamma^{\sigma}\gamma^{\alpha}\gamma^{5})=-4i\epsilon^{\rho\beta\sigma\alpha}, with ϵ0123=+1\epsilon^{0123}=+1. We use the antisymmetry of the Levi-Civita symbol to eliminate the term ϵαβρσ(kC)ρ(kC)σ\epsilon^{\alpha\beta\rho\sigma}(k^{\prime}-C)_{\rho}(k^{\prime}-C)_{\sigma} in the numerator. Since we are interested in the contribution to the effective action including the product of two electromagnetic tensors without additional derivatives, we calculate the integral Eq. (18) only to first order in pαp_{\alpha}. Finally, and assuming that mμνm^{\mu}{}_{\nu} is invertible, the identity mμmνβmκαϵβαρσσ=(detm)(m1)ρϵμνλκλm^{\mu}{}_{\beta}\,m^{\nu}{}_{\alpha}\,m^{\kappa}{}_{\sigma}\epsilon^{\beta\alpha\rho\sigma}=(\det m)\,(m^{-1})^{\rho}{}_{\lambda}\epsilon^{\mu\nu\lambda\kappa} yields the further simplification

ΠA,χμν(p)=2χe2(detm)(m1)ρϵμνλκλpκIρ(C),\displaystyle\Pi_{A,\,\chi}^{\mu\nu}(p)=-2{\chi}e^{2}(\det m)(m^{-1})^{\rho}{}_{\lambda}\epsilon^{\mu\nu\lambda\kappa}p_{\kappa}\,I_{\rho}(C), (20)

where we have defined the integral

Iρ(C)=d4k(2π)4gρ(k0,𝒌),gρ(k0,𝒌)=(kC)ρ[(kC)2]2,\displaystyle I_{\rho}(C)=\int\frac{d^{4}k}{\left(2\pi\right)^{4}}\,g_{\rho}(k_{0},{\mbox{\boldmath$k$}})\,,\quad g_{\rho}(k_{0},{\mbox{\boldmath$k$}})=\frac{\left(k^{\prime}-C\right)_{\rho}}{\left[\left(k^{\prime}-C\right)^{2}\right]^{2}}, (21)

and we recall that kμ=kαmαμ{k^{\prime}}_{\mu}={k}_{\alpha}m^{\alpha}{}_{\mu}. Previous to regularization, the above expression is our final result for the vacuum polarization tensor in Minkowski spacetime, which can be evaluated for the left- and right-handed fermions according to Eq. (20) with the assignments in Eq. (19). The result (20) holds for arbitrary LIV terms cμ,νdμ,νaμc^{\mu}{}_{\nu},\,d^{\mu}{}_{\nu},\,a_{\mu} and bμb_{\mu} as long as these produce invertible matrices lμνl^{\mu}{}_{\nu} and rμνr^{\mu}{}_{\nu}.

III The regularization

As a first step in the inclusion of thermal field theory methods in the description of the radiative corrections to the fermionic action (1) under the new conditions (3), here we shall consider the case of a non-zero chemical potential μ\mu but still remain in the zero-temperature limit. The inclusion of both will be published elsewhere. To this end we adopt the finite temperature approach in the imaginary time formulation KAPUSTA , which we recover after the substitution BERNARD ; DITTRICH ; DAS

d4k(2π)4+iTn=d3𝒌(2π)3,\displaystyle\int\frac{d^{4}k}{(2\pi)^{4}}\ \ \rightarrow\ \ +iT\sum_{n=-\infty}^{\infty}\int\frac{d^{3}{\mbox{\boldmath$k$}}}{(2\pi)^{3}}, (22)

with k0k0=iωn+Λk_{0}\to k_{0}=i\omega_{n}+\Lambda.

In order to have a specific system which determines the choice of the finite but undetermined parameters, together with a better physical understanding of the subsequent steps, we introduce the Hamiltonian

Hχ(𝒑)=𝒗χ(𝒑+χ𝒃~)χb~0+χvF𝝈(𝒑+χ𝒃~),\displaystyle H_{\chi}(\mbox{\boldmath$p$})={\mbox{\boldmath$v$}}_{\chi}\cdot({\mbox{\boldmath$p$}}+\chi{\mbox{\boldmath$\tilde{b}$}})-\chi{\tilde{b}}_{0}+\chi v_{F}{\mbox{\boldmath$\sigma$}}\cdot({\mbox{\boldmath$p$}}+\chi{\mbox{\boldmath$\tilde{b}$}}), (23)

borrowed from condensed matter, which describes a Weyl semimetal with two linear 3D band crossings of chirality χ=±1\chi=\pm 1 close to ±𝒃~\pm{\mbox{\boldmath$\tilde{b}$}} in momentum space and ±b~0\pm{\tilde{b}}_{0} in energy. Here vFv_{F} is the isotropic Fermi velocity at each band crossing, 𝝈=(σx,σy,σz){\mbox{\boldmath$\sigma$}}=(\sigma_{x},\sigma_{y},\sigma_{z}) is the triplet of spin-1/21/2 Pauli matrices, 𝒗χ{\mbox{\boldmath$v$}}_{\chi} is the tilting parameter and 𝒑p is the momentum. In the Weyl basis for the gamma matrices

γ0=(0σ0σ00),γi=(0σiσi0),γ5=(σ000σ0),\gamma^{0}=\begin{pmatrix}0&\sigma_{0}\\ \sigma_{0}&0\end{pmatrix},\hskip 14.22636pt\gamma^{i}=\begin{pmatrix}0&\sigma^{i}\\ -\sigma^{i}&0\end{pmatrix},\hskip 14.22636pt\gamma^{5}=\begin{pmatrix}-\sigma_{0}&0\\ 0&\sigma_{0}\end{pmatrix}, (24)

both 2×22\times 2 Hamiltonians in Eq. (23) can be embedded in the action (1) with the choices c0=01,c0=id0=id0=00c^{0}{}_{0}=1,c^{0}{}_{i}=d^{0}{}_{i}=d^{0}{}_{0}=0, i.e. Γ0=γ0\Gamma^{0}=\gamma^{0} together with Γi\Gamma^{i} and MM being determined by the parameters in (23). For simplicity, the anisotropy in the Fermi velocities is not included in (23) but enters naturally in our final result through the coefficients cijc^{i}{}_{j} and dijd^{i}{}_{j}. The Hamiltonian (23) is realized in terms of chiral fermions whose dispersion relation behaves linearly, with band crossing points localized both in momentum and energy, as schematically indicated in the Fig. 1(a). As usual, the position of the chemical potential determines most of the transport properties of a given material. In a metal for instance, it has to be measured from the gap closing, since it represents the filling of either the conduction or the valence bands. This suggests that in WSMs, the transport properties will depend on the band filling, i.e. on the chemical potential as measured from the node of the cones. Therefore, in order to apply the finite temperature approach (22) to WSMs, the parameter Λ\Lambda in Eq. (22) has to be understood as the chemical potential measured from the band-crossing points, as shown in Fig. 1(a). To be precise, Λ=μEχ(𝒑χ)\Lambda=\mu-E_{\chi}({\mbox{\boldmath$p$}}_{\chi}), where 𝒑χ{\mbox{\boldmath$p$}}_{\chi} is the location in momentum of the node with chirality χ\chi, and Eχ(𝒑χ)E_{\chi}({\mbox{\boldmath$p$}}_{\chi}) the corresponding energy. Both 𝒑χ{\mbox{\boldmath$p$}}_{\chi} and Eχ(𝒑χ)E_{\chi}({\mbox{\boldmath$p$}}_{\chi}) will be determined later for the problem at hand. In the simplified model (3), it is clear that 𝒑χ=χ𝒃~{\mbox{\boldmath$p$}}_{\chi}=\chi{\mbox{\boldmath$\tilde{b}$}} and Eχ(𝒑χ)=χb~0E_{\chi}({\mbox{\boldmath$p$}}_{\chi})=-\chi{\tilde{b}}_{0}.

The sum in Eq. (22) is over the Matsubara frequencies ωn=(2n+1)πT\omega_{n}=(2n+1)\pi T required to produce anti-periodic boundary conditions for the fermions KAPUSTA . Next we focus in Eq. (21) and we make use of the additional relation KAPUSTA

limT0Tn=gρ(k0=iωn+Λ,𝒌)\displaystyle\lim_{T\to 0}T\sum_{n=-\infty}^{\infty}g_{\rho}(k_{0}=i\omega_{n}+{\Lambda},{\mbox{\boldmath$k$}}) =12πiii𝑑k0gρ(k0,𝒌)+\displaystyle=\frac{1}{2\pi i}\int_{-i\infty}^{i\infty}dk_{0}\,g_{\rho}(k_{0},{\mbox{\boldmath$k$}})\,+
+12πiΩ𝑑k0gρ(k0,𝒌),\displaystyle\phantom{=}+\frac{1}{2\pi i}\oint_{\Omega}dk_{0}\,g_{\rho}(k_{0},{\mbox{\boldmath$k$}}), (25)

where the contour Ω\Omega is shown in the Fig. 1(b).

Refer to caption
Figure 1: (a) The effective chemical potential. (b) The contour of integration in the k0k_{0}-plane.

III.1 The μ\mu-independent contribution

The first term in the right hand side of Eq. (25) reduces to the standard zero-temperature, zero-chemical potential contribution, which has been previously discussed in the literature as extensively reported in section I. Going back to Eq. (20), this corresponds to the direct evaluation of the integral Iρ(C)I_{\rho}(C) after the change of integration variables kμ=kνmνμk^{\prime}_{\mu}=k_{\nu}m^{\nu}{}_{\mu}. Since the only vector at our disposal is CμC_{\mu} we have

Iρ(1)(C)=i(detm)N¯Cρ,N¯=1C2[d4k(2π)4(kC)C[(kC)2]2]E,I^{(1)}_{\rho}(C)=\frac{i}{(\det m)}\,{\bar{N}}C_{\rho},\quad{\bar{N}}=\frac{1}{C^{2}}\left[\int\frac{d^{4}k^{\prime}}{\left(2\pi\right)^{4}}\;\frac{\left(k^{\prime}-C\right)\cdot C}{\left[\left(k^{\prime}-C\right)^{2}\right]^{2}}\right]_{E}, (26)

where the integral inside the square brackets is in Euclidean space and the factor +i+i comes from the Wick rotation. The factor N¯{\bar{N}} is regularization dependent and could only be a function of the magnitude of the four-vector CμC_{\mu}. However, a change of scale CσλCσC_{\sigma}\to\lambda C_{\sigma} followed by an additional change of variables kμ′′=λkμk^{\prime\prime}_{\mu}=\lambda k^{\prime}_{\mu} shows that N¯{\bar{N}} is just a numerical factor, independent of CμC_{\mu}. Therefore, N¯{\bar{N}} is the same for both left- and right-handed fermions, i.e. N¯L=N¯R=N¯{\bar{N}}_{L}={\bar{N}}_{R}={\bar{N}}. In this way, the total contribution to the vacuum polarization in this case is summarized in the four-vector

λ(1)=4π2N¯[CLρ(l1)ρλCRρ(r1)ρ]λ,\displaystyle{\cal B}^{(1)}_{\lambda}=-4\pi^{2}{\bar{N}}\,\Big{[}C_{L\rho}\,(l^{-1})^{\rho}{}_{\lambda}-C_{R\rho}\,(r^{-1})^{\rho}{}_{\lambda}\Big{]}, (27)

according to Eq. (10). As shown previously in the literature the factor N¯{\bar{N}} is finite but undetermined. In the case of WSMs, its dependence upon the regularization procedure has been studied in Ref. Goswami and the final choice is made by selecting the anomalous Hall conductivity σxy=e2b~z/(2π2\sigma_{xy}=-e^{2}{\tilde{b}}_{z}/(2\pi^{2}) as the physical quantity to be reproduced in the zero-tilting limit of the Hamiltonian (23) Burkov ; Goswami ; BURKOVBALENTS . To this end it is necessary to take a cut-off in the direction of the spatial component of CC, with the result N¯=1/(8π2){\bar{N}}=-1/(8\pi^{2}). The suitability of this choice will be demonstrated later by comparing our field-theoretic results with those obtained from a semiclassical Boltzmann approach, which has proven to be successful in the study the transport properties of WSMs.

III.2 The μ\mu-dependent contribution

Next we consider the second term in the right-hand side of Eq. (25) and calculate

Iρ(2)(C)=d3𝒌(2π)312πiΩ𝑑k0gρ(k0,𝒌).\displaystyle I^{(2)}_{\rho}(C)=\int\frac{d^{3}{\mbox{\boldmath$k$}}}{(2\pi)^{3}}\,\frac{1}{2\pi i}\,\oint_{\Omega}dk_{0}\,g_{\rho}(k_{0},{\mbox{\boldmath$k$}}). (28)

Let us start by finding the double poles of the integrand in the k0k_{0}-plane. They are located at the two points k0±=C0±|𝒌𝑪|k^{\prime}_{0\pm}=C_{0}\pm|{\mbox{\boldmath$k$}}^{\prime}-{\mbox{\boldmath$C$}}|, where we recall the relations k0=kμmμ0k^{\prime}_{0}=k_{\mu}m^{\mu}{}_{0} and ki=k0m0+ikjmjik^{\prime}_{i}=k_{0}m^{0}{}_{i}+k_{j}m^{j}{}_{i}. In the general case the solution for k0k_{0} involves a rather cumbersome second order equation which will not be very illuminating for our purposes. In order to illustrate the full procedure we restrict ourselves to the simpler situation where m0=νδν0m^{0}{}_{\nu}=\delta^{0}_{\nu}. Under this assumption the poles are located at

k0ξ=kjmj+0C0+ξ|𝒌𝑪|,ξ=±1,kj=kimi,j\displaystyle k_{0\xi}=-k_{j}m^{j}{}_{0}+C_{0}+\xi\,|{\mbox{\boldmath$k$}}^{\prime}-{\mbox{\boldmath$C$}}|,\quad\xi=\pm 1,\quad k^{\prime}_{j}=k_{i}\,m^{i}{}_{j}, (29)

which corresponds to the dispersion relation of the model, wherefrom we can determine the position in momentum and energy of the Dirac/Weyl node. Clearly, the band touching point (i.e. the node) occurs at 𝒌=𝑪{\mbox{\boldmath$k$}}^{\prime}={\mbox{\boldmath$C$}}, and the corresponding energy is E0k0ξ(𝒌=𝑪)E_{0}\equiv k_{0\xi}({\mbox{\boldmath$k$}}^{\prime}={\mbox{\boldmath$C$}}), which can be explicitly expressed as

E0=𝒱jCj+C0=𝓥𝑪+C0,𝒱j=(m1)jmii.0E_{0}=-\mathcal{V}^{j}C_{j}+C_{0}=\mbox{\boldmath$\mathcal{V}$}\cdot{\mbox{\boldmath$C$}}+C_{0},\quad\mathcal{V}^{j}=(m^{-1})^{j}{}_{i}\,m^{i}{}_{0}. (30)

Applying the residue theorem we obtain

12πiΩ𝑑k0gρ=ξ=±1Res(gρ,k0=k0ξ)H(k0ξ)H(Λk0ξ),\displaystyle\frac{1}{2\pi i}\oint_{\Omega}dk_{0}\,g_{\rho}=\sum_{\xi=\pm 1}\mathrm{Res}(g_{\rho},k_{0}=k_{0\xi})H(k_{0\xi})H(\Lambda-k_{0\xi}), (31)

where the Heaviside functions HH guarantee that the poles k0ξk_{0\xi} fall inside the contour Ω\Omega of Fig. 1, i.e. 0<k0ξ<Λ0<k_{0\xi}<\Lambda, where Λ=μE0\Lambda=\mu-E_{0}, with E0E_{0} given by Eq. (30). The residues are

Res(gρ,k0=k0ξ)=ξ14δρikiCi|𝒌𝑪|3.\mathrm{Res}(g_{\rho},k_{0}=k_{0\xi})=-\xi\frac{1}{4}\delta^{i}_{\rho}\frac{k^{\prime}_{i}-C_{i}}{|{\mbox{\boldmath$k$}}^{\prime}-{\mbox{\boldmath$C$}}|^{3}}. (32)

Substitution of the residues (32) in Eq. (28) determines Iρ(2)(C)I^{(2)}_{\rho}(C), and combining the result with Eq. (20) it follows that the vacuum polarization becomes

ΠA,χμν(p)\displaystyle{\Pi}_{A,\,\chi}^{\mu\nu}(p) =iχe22(detm)(m1)ipκλϵμνλκ\displaystyle=i\chi\,\frac{e^{2}}{2}(\det m)(m^{-1})^{i}{}_{\lambda}p_{\kappa}\epsilon^{\mu\nu\lambda\kappa}
×ξ=±1ξd3𝒌(2π)3kiCi|𝒌𝑪|3H(k0ξ)H(Λk0ξ),\displaystyle\phantom{=}\times\sum_{\xi=\pm 1}\xi\int\frac{d^{3}{\mbox{\boldmath$k$}}}{(2\pi)^{3}}\frac{k^{\prime}_{i}-C_{i}}{|{\mbox{\boldmath$k$}}^{\prime}-{\mbox{\boldmath$C$}}|^{3}}H(k_{0\xi})H(\Lambda-k_{0\xi}), (33)

which includes the factor +i+i arising from the Eq. (22). It is convenient to make the change of variables

ki′′=kiCi,d3𝒌′′=d3𝒌=det(m)d3𝒌\displaystyle k^{\prime\prime}_{i}=k^{\prime}_{i}-C_{i},\quad d^{3}{\mbox{\boldmath$k$}}^{\prime\prime}=d^{3}{\mbox{\boldmath$k$}}^{\prime}=\det(m)\,d^{3}{\mbox{\boldmath$k$}} (34)

where we used that det(mi)j=det(mμ)ν=det(m)\det(m^{i}{}_{j})=\det(m^{\mu}{}_{\nu})=\det(m) since m0=νδν0m^{0}{}_{\nu}=\delta^{0}_{\nu}. This yields

ΠA,χμν(p)=iχe22(m1)ipκλϵμνλκ(Ii+Ii)\displaystyle{\Pi}^{\mu\nu}_{A,\,\chi}(p)=i\chi\frac{e^{2}}{2}(m^{-1})^{i}{}_{\lambda}\,p_{\kappa}\,\epsilon^{\mu\nu\lambda\kappa}\,\big{(}I_{i}^{+}-I_{i}^{-}\big{)} (35)

with

Iiξ=d3𝒌′′(2π)3ki′′|𝒌′′|3H(k0ξ)H(Λk0ξ).I^{\xi}_{i}=\int\frac{d^{3}{\mbox{\boldmath$k$}}^{\prime\prime}}{(2\pi)^{3}}\frac{k^{\prime\prime}_{i}}{|{\mbox{\boldmath$k$}}^{\prime\prime}|^{3}}\,H(k_{0\xi})\,H(\Lambda-k_{0\xi}). (36)

In the new double-primed variables k′′k^{\prime\prime} the constraints 0<k0ξ<Λ0<k_{0\xi}<\Lambda read

0<(kj′′+Cj)(m1)jmii+0C0+ξ|𝒌′′|<Λ.\displaystyle 0<-(k^{\prime\prime}_{j}+C_{j})(m^{-1})^{j}{}_{i}\,m^{i}{}_{0}+C_{0}+\xi\,|{\mbox{\boldmath$k$}}^{\prime\prime}|<\Lambda. (37)

From the identifications in Eq. (30) these constrains become

0<𝓥𝒌′′+ξ|𝒌′′|+E0<Λ.\displaystyle 0<\mbox{\boldmath$\mathcal{V}$}\cdot{\mbox{\boldmath$k$}}^{\prime\prime}+\xi\,|{\mbox{\boldmath$k$}}^{\prime\prime}|+E_{0}<\Lambda. (38)

It is convenient to perform the 3-momentum integration in spherical coordinates. Hence we present the condition (38) as

0<|𝒌′′|(|𝓥|cosθ+ξ)+E0<Λ,\displaystyle 0<|{\mbox{\boldmath$k$}}^{\prime\prime}|\,\big{(}|\mbox{\boldmath$\mathcal{V}$}|\cos\theta+\xi\big{)}+E_{0}<\Lambda, (39)

Writing IiξI_{i}^{\xi} as a linear combination of the vectors 𝒱i{\cal V}_{i} and CiC_{i}, one can show that the projection upon CiC_{i} is zero, so that

Iiξ=Nξ𝒱i,\displaystyle I^{\xi}_{i}=N^{\xi}\,\mathcal{V}_{i}, (40)

where NξN^{\xi} depends only on (|𝓥|,E0,Λ)(|\mbox{\boldmath$\mathcal{V}$}|,E_{0},\Lambda) as

Nξ=1(2π)2|𝓥|0π𝑑θsinθcosθ0d|𝒌′′|H(k0ξ)H(Λk0ξ).\displaystyle N^{\xi}=\frac{1}{(2\pi)^{2}|\mbox{\boldmath$\mathcal{V}$}|}\int_{0}^{\pi}\!\!d\theta\sin\theta\cos\theta\int_{0}^{\infty}\!\!d|{\mbox{\boldmath$k$}}^{\prime\prime}|\,H(k_{0\xi})H(\Lambda-k_{0\xi}). (41)

Note that for |𝓥|1|\mbox{\boldmath$\mathcal{V}$}|\geq 1 there exists an angular direction for which |𝓥|cosθ+ξ|\mbox{\boldmath$\mathcal{V}$}|\cos\theta+\xi is zero whereby the integral over |𝒌′′||{\mbox{\boldmath$k$}}^{\prime\prime}| diverges and which would imply the need of a nontrivial renormalization procedure. Thus, from now on we take the simpler convergent case |𝓥|<1|\mbox{\boldmath$\mathcal{V}$}|<1 (relevant for type-I WSMs) which yields |𝓥|cosθ+ξ|\mbox{\boldmath$\mathcal{V}$}|\cos\theta+\xi with definite sign: positive (negative) for ξ=+1(ξ=1)\xi=+1(\xi=-1). This in turn fixes the inequalities indicated in Eq. (39) as

ξ=+1\displaystyle\xi=+1 :E0|𝓥|cosθ+1<|𝒌′′|<ΛE0|𝓥|cosθ+1\displaystyle:\quad\frac{-E_{0}}{|\mbox{\boldmath$\mathcal{V}$}|\cos\theta+1}<|{\mbox{\boldmath$k$}}^{\prime\prime}|<\frac{\Lambda-E_{0}}{|\mbox{\boldmath$\mathcal{V}$}|\cos\theta+1} (42)
ξ=1\displaystyle\xi=-1 :ΛE0|𝓥|cosθ1<|𝒌′′|<E0|𝓥|cosθ1\displaystyle:\quad\frac{\Lambda-E_{0}}{|\mbox{\boldmath$\mathcal{V}$}|\cos\theta-1}<|{\mbox{\boldmath$k$}}^{\prime\prime}|<\frac{-E_{0}}{|\mbox{\boldmath$\mathcal{V}$}|\cos\theta-1} (43)

The above equations demand us to distinguish the signs of E0E_{0} and of ΛE0\Lambda-E_{0}, since we must ensure that |𝒌′′|0|{\mbox{\boldmath$k$}}^{\prime\prime}|\geq 0. This leaves us with four cases according to the combinations ΛE00{\Lambda-E_{0}}\gtrless 0 and E00E_{0}\gtrless 0. Here we make use of the assumption that Λ>0\Lambda>0, as it constrains the possibilities of these four cases. Under this condition a detailed analysis yields the global result

Ii+Ii=Λπ2Nχ𝒱i,Nχ=12|𝓥|3(|𝓥|arctanh(|𝓥|)).\displaystyle I^{+}_{i}-I^{-}_{i}=\frac{\Lambda}{\pi^{2}}N_{\chi}{\mathcal{V}}_{i},\quad N_{\chi}=\frac{1}{2|\mbox{\boldmath$\mathcal{V}$}|^{3}}\Big{(}|\mbox{\boldmath$\mathcal{V}$}|-\mathrm{arctanh}(|\mbox{\boldmath$\mathcal{V}$}|)\Big{)}. (44)

The resulting contribution to the vacuum polarization is

ΠAμν=ie22π2pκϵμνλκ[ΛRNR𝒱Ri(r1)iλΛLNL𝒱Li(l1)i]λ,\Pi^{\mu\nu}_{A}=i\frac{e^{2}}{2\pi^{2}}p_{\kappa}\,\epsilon^{\mu\nu\lambda\kappa}\,\Big{[}\Lambda_{R}\,N_{R}\,{\mathcal{V}}_{Ri}\,(r^{-1})^{i}{}_{\lambda}-\Lambda_{L}\,N_{L}\,{\mathcal{V}}_{Li}\,(l^{-1})^{i}{}_{\lambda}\Big{]}, (45)

where Λ\Lambda has acquired a chirality dependence through the splitting of the zero excitation energy modes given by

E0L(R)=𝓥L(R)𝑪L(R)+C0L(R),ΛL(R)=μE0L(R).E_{0L(R)}=\mbox{\boldmath$\mathcal{V}$}_{L(R)}\cdot{\mbox{\boldmath$C$}_{L(R)}}+C_{0L(R)},\quad\Lambda_{L(R)}=\mu-E_{0L(R)}. (46)

We thus obtain the additional contribution

λ(2)=[ΛRNR𝒱Ri(r1)iλΛLNL𝒱Li(l1)i]λ\displaystyle{\cal B}^{(2)}_{\lambda}=-\Big{[}\Lambda_{R}\,N_{R}\,{\mathcal{V}}_{Ri}\,(r^{-1})^{i}{}_{\lambda}-\Lambda_{L}\,N_{L}\,{\mathcal{V}}_{Li}\,(l^{-1})^{i}{}_{\lambda}\Big{]} (47)

to the coupling Θ(x)\Theta(x) in Eq. (10).

Summarizing, the full effective electromagnetic action of the system described by the fermionic action (1), with the only restrictions c0=νδν0c^{0}{}_{\nu}=\delta^{0}_{\nu},   d0=ν0d^{0}{}_{\nu}=0, lμνl^{\mu}{}_{\nu} and rμνr^{\mu}{}_{\nu} invertible and |𝓥|<1|\mbox{\boldmath$\mathcal{V}$}|<1, is given by the action (4) with Θ(x)=xλ(λ(1)+λ(2))\Theta(x)=x^{\lambda}\big{(}{\cal B}^{(1)}_{\lambda}+{\cal B}^{(2)}_{\lambda}\big{)}. Notice that the contribution proportional to μ\mu is not regularization dependent. The coupling Θ\Theta is CPT and PT odd as reflected in Eqs. (6) with b~μμ{\tilde{b}}_{\mu}\rightarrow{\cal B}_{\mu} since i{\cal B}_{i} breaks T but not P, while 0{\cal B}_{0} does the opposite. Then, even if we start with only the CPT even part of the Lagrangian in Eq. (1) (the Γμ\Gamma^{\mu} term ) by setting aμ=bμ=0a_{\mu}=b_{\mu}=0 it is not surprising that we obtain a non-zero PT odd Θ\Theta because Γμ\Gamma^{\mu} already includes both PT even and odd contributions. In this particular case λ(1)=0,Λχ=μ{\cal B}^{(1)}_{\lambda}=0,\,\Lambda_{\chi}=\mu, but λ(2)0{\cal B}^{(2)}_{\lambda}\neq 0 because 𝓥R{\mbox{\boldmath$\cal V$}}_{R} and 𝓥L{\mbox{\boldmath$\cal V$}}_{L} remain arbitrary. In other words, the source of the PT (CPT) odd effective electromagnetic action here is the γ5\gamma^{5} PT odd contribution in Γμ\Gamma^{\mu}.

One particularly interesting and simple system takes place when we take cj=iδjic^{j}{}_{i}=\delta^{j}{}_{i} and dj=i0d^{j}{}_{i}=0, which corresponds to the case of arbitrary tilting 𝑽R\mbox{\boldmath$V$}_{R} and 𝑽L\mbox{\boldmath$V$}_{L}, but with equal isotropic Fermi velocity vF=1v_{F}=1 at each node. That is, we take

CRν\displaystyle C_{R\nu} =aνbν,CLν=aν+bν,\displaystyle=a_{\nu}-b_{\nu},\qquad\qquad\quad C_{L\nu}=a_{\nu}+b_{\nu}, (48)
rμν\displaystyle r^{\mu}{}_{\nu} =δμ+νVRiδμδ0i,νlμ=νδμ+νVLiδμδ0i,ν\displaystyle=\delta^{\mu}{}_{\nu}+V^{i}_{R}\,\delta^{\mu}{}_{i}\,\delta^{0}{}_{\nu},\qquad l^{\mu}{}_{\nu}=\delta^{\mu}{}_{\nu}+V^{i}_{L}\,\delta^{\mu}{}_{i}\,\delta^{0}{}_{\nu}, (49)
(r1)μν\displaystyle(r^{-1})^{\mu}{}_{\nu} =δνμVRiδμδ0i,ν(l1)μ=νδνμVLiδμδ0i.ν\displaystyle=\delta^{\mu}_{\nu}-V^{i}_{R}\delta^{\mu}{}_{i}\delta^{0}{}_{\nu},\qquad(l^{-1})^{\mu}{}_{\nu}=\delta^{\mu}_{\nu}-V^{i}_{L}\delta^{\mu}{}_{i}\delta^{0}{}_{\nu}. (50)

Under these conditions 𝒱i=Vi{\cal V}^{i}=V^{i}. Putting together the contributions for λ=λ(1)+λ(2){\cal B}_{\lambda}={\cal B}^{(1)}_{\lambda}+{\cal B}^{(2)}_{\lambda} we obtain

0\displaystyle{\cal B}_{0} =b0+𝒃𝑼++𝒂𝑼χ=±1χΛχ2|𝑽χ|(|𝑽χ|arctanh(|𝑽χ|)),\displaystyle=b_{0}+{\mbox{\boldmath$b$}}\cdot{\mbox{\boldmath$U$}}_{+}+{\mbox{\boldmath$a$}}\cdot{\mbox{\boldmath$U$}}_{-}-\sum_{\chi=\pm 1}\frac{\chi\Lambda_{\chi}}{2|{\mbox{\boldmath$V$}}_{\chi}|}\,\Big{(}|{\mbox{\boldmath$V$}}_{\chi}|-\mathrm{arctanh}(|{\mbox{\boldmath$V$}}_{\chi}|)\Big{)},
i\displaystyle{\cal B}^{i} =biχ=±1(Vχ)iχΛχ2|𝑽χ|3(|𝑽χ|arctanh(|𝑽χ|)),\displaystyle=b^{i}-\sum_{\chi=\pm 1}(V_{\chi})^{i}\,\frac{\chi\Lambda_{\chi}}{2|{\mbox{\boldmath$V$}}_{\chi}|^{3}}\,\Big{(}|{\mbox{\boldmath$V$}}_{\chi}|-\mathrm{arctanh}(|{\mbox{\boldmath$V$}}_{\chi}|)\Big{)}, (51)

with 𝑼±=12(𝑽±𝑽+){\mbox{\boldmath$U$}}_{\pm}=\frac{1}{2}({\mbox{\boldmath$V$}}_{-}\pm{\mbox{\boldmath$V$}}_{+}) and Λχ=μE0χ\Lambda_{\chi}=\mu-E_{0\chi} according to Eq. (46). We recall that χ=1\chi=1(χ=1\chi=-1) denote the R(L)R(L) contributions. Consistently with the property that the axial anomaly is insensitive to LIV modifications LUAGA ; FIDEL ; SALVIO ; Scarpelli , our results (51) show that the Pontryagin density remains unchanged, and that the additional LIV terms in Eq. (2) which defines the fermionic action (1), as well as the chemical potential, modify only the Θ\Theta coupling.

In order to check the consistency of our results with a condensed matter approach, we first establish the conditions under which our general model reduces to that of a WSM as described by the Hamiltonian (23). Indeed, the equivalence is achieved by setting 𝑽χ=𝒗χ,𝑼χ=𝒖χ{\mbox{\boldmath$V$}}_{\chi}={\mbox{\boldmath$v$}}_{\chi},\,{\mbox{\boldmath$U$}}_{\chi}={\mbox{\boldmath$u$}}_{\chi}, and

𝒂=0,𝒃=𝒃~,a0=𝒃~𝒖,b0=b~0𝒃~𝒖+{\mbox{\boldmath$a$}}=0,\quad{\mbox{\boldmath$b$}}={\mbox{\boldmath$\tilde{b}$}},\quad a_{0}=-{\mbox{\boldmath$\tilde{b}$}}\cdot{\mbox{\boldmath$u$}}_{-},\quad b_{0}={\tilde{b}}_{0}-{\mbox{\boldmath$\tilde{b}$}}\cdot{\mbox{\boldmath$u$}}_{+} (52)

such that Λχ=μ+χb~0\Lambda_{\chi}=\mu+\chi{\tilde{b}}_{0}. In particular, for a type-I Weyl semimetal (i.e. with vχ<vF=1v_{\chi}<v_{F}=1) with the tilting 𝒗χ{\mbox{\boldmath$v$}}_{\chi} parallel to 𝒃~\tilde{b}, the semiclassical Boltzmann approach Kubo2 leads to the same effective action (4) with the Weyl node separation shifted by ENPREP

b~b~χ=±1χΛχ2|vχ|2[vχarctanh(vχ)],\displaystyle{{\tilde{b}}}\;\to\;{\tilde{b}}\,-\sum_{\chi=\pm 1}\frac{\chi\Lambda_{\chi}}{2|{v}_{\chi}|^{2}}\,\Big{[}v_{\chi}-{\rm arctanh}(v_{\chi})\Big{]}, (53)

where b~=|𝒃~|{\tilde{b}}=|\mbox{\boldmath$\tilde{b}$}| and Λχ=μ+χb~0>0\Lambda_{\chi}=\mu\,+\chi{\tilde{b}}_{0}>0 is the chemical potential measured from the nodal point. Clearly, the spatial components in Eq. (51) successfully simplify to the result of Eq. (53).

IV Summary and conclusions

We extend the vacuum polarization method of high energy physics, used in the calculation of the electromagnetic response of fermionic systems, to a large class of fermionic couplings included in the Standard Model Extension (SME) describing Lorentz symmetry violations Kostelecky0 . Emphasis is made in the CPT odd contribution to the effective action which exhibits remnants of the abelian chiral anomaly due to the appearance of the Pontryagin density ϵμνρσFμνFρσ\epsilon^{\mu\nu\rho\sigma}F_{\mu\nu}F_{\rho\sigma}. Our approach does not rest in a direct manipulation of chiral transformations which induce the chiral anomaly via the corresponding Jacobian Fujikawa . Rather, it is based in the standard perturbative expansion in powers of the electromagnetic potential of the determinant resulting from the integration of the fermions in the functional integral corresponding to the action (1) plus the specific choices (2) . This amounts to the calculation of the vacuum polarization tensor to second order in AμA_{\mu} starting from the exact modified fermion propagators obtained from the SME. A distinguishing feature of the method is the inclusion of corrections depending upon the chemical potential μ\mu, which enlarge the range of possible phenomenological applications. This is possible through a systematic application of thermal field theory methods KAPUSTA , which is exemplified here in the case of the zero temperature limit. The fermionic systems we consider exhibit the axial anomaly and the resulting electromagnetic actions are fully described by axion electrodynamics in the form of Eq. (4). All corrections induced by the different parameters in the SME enter only in the coupling Θ(x)=2λxλ\Theta(x)=2{\cal B}_{\lambda}x^{\lambda} to the Pontryagin density, as shown in our general expressions (27) and (47). This is consistent with the property that the axial anomaly is insensitive to the Lorentz invariance violating parameters of the SME LUAGA ; FIDEL ; SALVIO ; Scarpelli . Along the text we have insisted in the advantage of extending these high energy physics methods to condensed matter physics. In fact we envisage interesting applications in the realm of topological quantum matter, where a rich phenomenology in the electromagnetic response can be explored. Our general results can be applied to type-I Weyl semimetals (WSM) with arbitrary tilting and anisotropies. A simpler system results from the restriction to a WSM with arbitrary tilting but equal isotropic Fermi velocity at each node, described by the Hamiltonian (23). In this setting, our result (51) for the vector contribution i{\cal B}^{i} to the Θ\Theta coupling has been validated by a direct calculation of the conductivity using the Boltzmann semiclassical approach, as shown in Eq. (53). The calculation and proper regularization of the corrections for the case |𝓥|1|\mbox{\boldmath$\mathcal{V}$}|\geq 1 (relevant for type-II WSM) is left unsettled for future research. Pending also is the introduction of a chiral chemical potential which is allowed by the chiral structure of the Hamiltonian describing a WSM with two band-crossings, i.e. each node can be held at different chemical potential. We find appealing also to pursue the comparison between the quantum field theory approach and the semiclassical Boltzmann formalism. In particular, it would be interesting to elucidate the role of the Berry phase in the first strategy ENPREP . Finally, the recent calculation of the one-loop Heisenberg-Euler effective action for zero chemical potential in two of the most studied minimal Lorentz-violating extensions of QED PETROV , motivates the challenge of extending the thermal field theory approach to the calculation of the CPT even contributions to the effective electromagnetic action arising from the fermionic sector of the SME considered in Eq. (2).

Acknowledgements

L.F.U. and A.M.-R. acknowledge support from the project CONACYT (México) # CF-428214. A.M.-R. has been partially supported by DGAPA-UNAM Projects # IA101320 and # IA102722. L.F.U. and A.G.A. were supported in part by Project DGAPA-UNAM # 103319.

References

  • (1) D. Colladay and V.A. Kostelecký, CPT violation and the standard model, Phys. Rev. D 55 (1997) 6760.
  • (2) D. Colladay and V.A. Kostelecký, Lorentz-violating extension of the standard model, Phys. Rev. D 58 (1998) 116002.
  • (3) S. Coleman and S. L. Glashow, High-energy tests of Lorentz invariance, Phys. Rev. D 59(1999) 116008.
  • (4) R. Jackiw and V. A. Kostelecký,Radiatively Induced Lorentz and CPT Violation in Electrodynamics, Phys. Rev. Lett. 82 (1999) 3572.
  • (5) M. Pérez-Victoria, Exact Calculation of the Radiatively Induced Lorentz and CPT Violation in QED, Phys. Rev. Lett. 83(1999)2518.
  • (6) J. M. Chung, Radiatively-Induced Lorentz and CPT violating Chern-Simons Terms in QED, Phys. Lett. B 461(1999) 138.
  • (7) J. M. Chung and P. Oh, Lorentz and CPT Violating Chern-Simons Term in the Derivative Expansion of QED, Phys. Rev. D 60 (1999) 067702.
  • (8) W. F. Chen, Understanding radiatively induced Lorentz-CPT violation in differential regularization, Phys. Rev. D 60(1999) 085007.
  • (9) G. Bonneau, Regularisation : many recipes, but a unique principle : Ward identities and Normalisation conditions. The case of CPT violation in QED, Nucl. Phys. B 593(2001) 398.
  • (10) M. Chaichian, W. F. Chen, and R. G. Felipe, Radiatively induced Lorentz and CPT violation in Schwinger constant field approximation, Phys. Lett. B 503 (2001) 215.
  • (11) M. Pérez-Victoria, Physical (ir)relevance of ambiguities to Lorentz and CPT violation in QED, J. High Energy Phys. 04 (2001) 032.
  • (12) O. A. Battistel and G. Dallabona, Role of Ambiguities and Gauge Invariance in the Calculation of the Radiatively Induced Chern-Simons Shift in Extended Q.E.D., Nucl. Phys. B 610 (2001)316.
  • (13) A. A. Andrianov, P. Giacconi, and R. Soldati, Lorentz and CPT violations form Chern-Simons modifications of QED, J. High Energy Phys. 02 (2002) 030.
  • (14) B. Altschul, Failure of gauge invariance in the nonperturbative formulation of massless Lorentz-violating QED, Phys. Rev. D 69 (2004) 125009.
  • (15) B. Altschul, Gauge invariance and the Pauli-Villars regulator in Lorentz- and CPT-violating electrodynamics, Phys. Rev. D 70 (2004) 101701(R).
  • (16) J. Alfaro, A. A. Andrianov, M. Cambiaso, P. Giacconi, R. Soldati, Bare and Induced Lorentz & CPT Invariance Violations in QED, Int. J. Mod. Phys. A 25(2010)3271.
  • (17) R. Jackiw, When radiative corrections are finite but undetermined, Int. J. Mod. Phys B 14 (2000) 2011.
  • (18) B. Altschul, There is no ambiguity in the radiatively induced gravitational Chern-Simons term, Phys. Rev. D 99(2019)125009.
  • (19) V. A. Kostelecký and N. Russell, Data tables for Lorentz and CPT violation, Rev. Mod. Phys. 83 (2011) 11 .
  • (20) Grushin, A.G. Consequences of a condensed matter realization of Lorentz violating QED in Weyl semi-metals. Phys. Rev. D 86 (2012) 045001.
  • (21) A.A. Zyuzin, A.A Burkov, Topological response in Weyl semimetals and the chiral anomaly, Phys. Rev. B, 86 (2012) 115133.
  • (22) P. Goswami, S. Terawi, Axionic field theory of (3 + 1)-dimensional Weyl semimetals, Phys. Rev. B 88 (2013) 245107.
  • (23) A. Sekine and K. Nomura, Axion electrodynamics in topological materials, J. Appl. Phys. 129 (2021) 141101.
  • (24) M. N. Chernodub, Y. Ferreiros, A. G. Grushin, K. Landsteiner, María A. H. Vozmediano, Thermal transport, geometry, and anomalies, arXiv:2110.05471.
  • (25) V. A. Kostelecký, R. Lehnert, N. McGinnis, M. Schreck and B. Seradjeh, Lorentz violation in Dirac and Weyl semimetals, arXiv:2112.14293.
  • (26) G. E. Volovik, On induced CPT-odd Chern-Simons terms in 3+13+1 effective action, Sov. Phys. JETP Lett. 70(1999)1.
  • (27) M. M. Vazifeh and M. Franz, Electromagnetic response of Weyl Semimetals, Phys. Rev. Lett. 111 (2013) 027201.
  • (28) A. A. Burkov and L. Balents, Weyl Semimetal in a Topological Insulator Multilayer, Phys. Rev. Lett. 107(2011)127205.
  • (29) K. Fujikawa, H. Suzuki, Path Integrals and Quantum Anomalies, Clarendon Press, Oxford, 2004, pp. 65, 92, 149.
  • (30) R.A. Bertlmann, Anomalies in Quantum Field Theory, Clarendon Press, Oxford, 1996.
  • (31) J. M. Chung, Lorentz- and CPT-violating Chern-Simons term in the functional integral formalism, Phys. Rev. D 60(1999) 127901.
  • (32) A. Gómez, L. Urrutia, The Axial Anomaly in Lorentz Violating Theories: Towards the Electromagnetic Response of Weakly Tilted Weyl Semimetals, Symmetry 13 (2021) 1181.
  • (33) P. Arias, H. Falomir, J. Gamboa, F. Mendez, F. Schaposnik, Chiral anomaly beyond Lorentz invariance, Phys. Rev. D, 76 (2007) 025019.
  • (34) A. Salvio, Relaxing Lorentz invariance in general perturbative anomalies, Phys. Rev. D 78 (2008) 085023.
  • (35) A.P.B. Scarpelli, T. Mariz, J.R. Nascimento, A.Y. Petrov, On the anomalies in Lorentz-breaking theories, Int. J. Mod. Phys. A, 31 (2016) 1650063.
  • (36) D. Ebert, V. Ch. Zhukovsky and A. S. Razumovsky, Chern-Simons like term generation in an extended model of QED under external conditions, Phys. Rev. D 70(2004) 025003.
  • (37) F. A. Brito, L. S. Grigorio, M. S. Guimaraes, E. Passos, and C. Wotzasek, Induced Chern-Simons-like action in Lorentz-violating massless QED, Phys. Rev. D 78 (2008) 125023.
  • (38) T. Mariz, J. R. Nascimento, E. Passos, R. F. Ribeiro and F. A. Brito, A remark on Lorentz violation at finite temperature, JHEP 10 (2005) 019
  • (39) J. R. Nascimento, E. Passos, A.Yu. Petrov and F. A. Brito, Lorentz-CPT violation, radiative corrections and finite temperature, JHEP 06 (2007) 016
  • (40) J. Leite and T. Mariz, Induced Lorentz-violating terms at finite temperature, EPL 99 (2012) 21003
  • (41) V. Ch. Zhukovsky, A. E. Lobanov, and E. M. Murchikova, Radiative effects in the standard model extension, Phys. Rev. D 73 (2006) 065016.
  • (42) M. E. Peskin, D. V. Schroeder, An introduction to Quantum Field Theory, Perseus Books, Cambridge, 1995.
  • (43) J. I. Kaputsa, C. Gale, Finite-temperature Field Theory: Principles and Applications, second ed., Cambridge University Press, Cambridge, 2006.
  • (44) C. W. Bernard, Feynman rules for gauge theories at finite temperature, Phys. Rev. D 9(1974) 3312.
  • (45) W. Ditrich, Effective Lagrangians at finite temperature, Phys. Rev. D 19(1979)2385.
  • (46) A. Das, Finite Temperature Field Theory, World Scientific, Singapore, 1997.
  • (47) R. Kubo, Statistical Mechanics, seventh ed., North-Holland Physics Publishing, Amsterdam, 1988, pp. 363-366.
  • (48) A. Gómez et al, The electromagnetic effective action for weakly tilted Weyl semimetals, (Unpublished results).
  • (49) A. F. Ferrari, J. Furtado, J. F. Assunção, T. Mariz and A. Yu. Petrov, One-loop calculations in Lorentz-breaking theories and proper-time method, EPL 136(2021)21002.