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

aainstitutetext: Departments of Physics, Keio University, 4-1-1 Hiyoshi, Kanagawa 223-8521, Japan bbinstitutetext: KEK Theory Center, Tsukuba 305-0801, Japaninstitutetext: Graduate University for Advanced Studies (Sokendai), Tsukuba 305-0801, Japaninstitutetext: RIKEN iTHEMS, RIKEN, Wako 351-0198, Japanccinstitutetext: Theoretical Research Division, Nishina Center, RIKEN, Wako, Saitama 351-0198, Japan

Second order chiral kinetic theory under gravity and antiparallel charge-energy flow

Tomoya Hayata b    Yoshimasa Hidaka c    and Kazuya Mameda hayata@keio.jp hidaka@post.kek.jp kazuya.mameda@riken.jp
Abstract

We derive the chiral kinetic theory under the presence of a gravitational Riemann curvature. It is well-known that in the chiral kinetic theory there inevitably appears a redundant ambiguous vector corresponding to the choice of the Lorentz frame. We reveal that on top of this conventional frame choosing vector, higher-order quantum correction to the chiral kinetic theory brings an additional degrees of freedom to specify the distribution function. Based on this framework, we derive new types of fermionic transport, that is, the charge current and energy-momentum tensor induced by the gravitational Riemann curvature. Such novel phenomena arise not only under genuine gravity but also in a (pseudo-)relativistic fluid, for which inhomogeneous vorticity or temperature are effectively represented by spacetime metric tensor. It is especially found that the charge and energy currents are antiparallelly induced by an inhomogeneous fluid vorticity (more generally, by the Ricci tensor R0i{R_{0}}^{i}), as a consequence of the spin-curvature coupling. We also briefly discuss possible applications to Weyl/Dirac semimetals and heavy-ion collision experiments.

preprint: KEK-TH-2279, J-PARC-TH-0234, RIKEN-QHP-488

1 Introduction and Summary

Transport phenomena are a pivotal subject in modern quantum field theories. Similar to external electromagnetic fields, (effective) background gravitational fields are intriguing sources to generate various currents. First, the most widely well-known example is fluid vorticity; the fluid velocity can be described by a metric tensor of the comoving frame with fluid. The corresponding transport phenomenon, i.e., the chiral vortical effect (CVE) Vilenkin:1979ui can be regarded as the gravitational counterpart of the chiral magnetic effect (CME) Vilenkin:1980fu ; Nielsen:1983rb ; Fukushima:2008xe , as is clear from the gravitoelectromagnetism Landsteiner:2011cp . The CVE is not only a theoretically interesting phenomenon in the sense that it is originated from quantum anomaly Son:2009tf ; Landsteiner:2011cp , but also an important experimental probe to study the rotation of quark-gluon plasmas created in relativistic heavy-ion collision experiments STAR:2017ckg . Second, the mechanical strain plays a role of an effective U(1)U(1) or axial U(1)U(1) magnetic field, and accordingly yields a charge current PhysRevLett.115.177202 ; Cortijo:2016wnf ; PhysRevX.6.041046 . Third, spacetime torsion is recently under active investigation, as it can bring novel currents, which are referred to as the chiral torsional effect Sumiyoshi:2015eda ; Khaidukov:2018oat ; Ferreiros:2018udw ; Imaki:2019ite ; Imaki:2020csc ; Gao:2020gcf .

In contrast to the aforementioned effects from the spacetime geometry, we do not fully understand the effect of the gravitational Riemann curvature in the quantum transport theory. Even at the classical level, however, its importance has been known; the trajectory of a spinning particle is modified by the Riemann curvature Mathisson:1937zz ; Dixon:1970zza ; Papapetrou:1951pa . In the context of quantum transport theory, this knowledge suggests that the Riemann curvature can be the trigger of a characteristic transport of the fermion chirality (or spin, more generally). In cosmological systems, such spacetime distortions may become dominant contributions to determine the fermionic transport rather than background electromagnetic fields. In laboratory environments, a fluid motion and temperature gradient can be described by effective gravities leading to non-vanishing Riemann curvatures. Therefore, the curvature-induced transport phenomena could be relevant in a wide range of physics from table-top experiments to the Universe.

For the nonequilibrium dynamics in the weak interaction regime, one of the promising theoretical implements would be the kinetic theory. In particular, the so-called chiral kinetic theory (CKT) Stephanov:2012ki ; Son:2012wh , which nicely reproduces the chiral anomaly, plays a pivotal role in the development of various studies of the chiral transport phenomena Son:2012zy ; Chen:2012ca ; Manuel:2013zaa ; Chen:2014cla ; Chen:2015gta ; Hidaka:2016yjf ; Hidaka:2017auj ; Mueller:2017lzw ; Mueller:2017arw ; Huang:2018wdl ; Carignano:2018gqt ; Dayi:2018xdy ; Liu:2018xip ; Lin:2019ytz ; Carignano:2019zsh in the context of heavy-ion collision, condensed matter and neutrino physics; although the kinetic theory is inapplicable to strongly-coupled quark-gluon plasmas, the early stage of heavy-ion collisions is described well by the Boltzmann transport theory Baym:1984np ; Mueller:1999pi ; Baier:2000sb . The CKT conventionally involves only the leading order quantum correction so that the anomalous aspects can be taken into account as the Berry curvature. However, the leading order CKT is insufficient to capture the gravitational curvature contributions to the transport coefficients, although the kinetic equation involves the spin-curvature coupling Liu:2018xip . As is readily expected, higher-order corrections make the theory much more complicated, and an intuitive deduction would not avail. This fact can be found from the equilibrium distribution function. The O()O(\hbar) contribution to enter the distribution is anticipated to be the spin-vorticity coupling, if we recall the conservation of the total angular momentum Chen:2015gta . On the other hand, this intuition is inapplicable to the O(2)O(\hbar^{2}) contribution, particularly, under a background gravitational field, as it is nontrivial to identify how the total angular momentum is modified at this order. Unlike the effective formalisms that relied on the Berry curvature, the derivation from quantum field theory works well against such a complication. In this case, the semiclassical (or weak coupling) dynamics is described by the Wigner transformation of the fermion propagator, which is a quantum-extended quantity of the classical distribution function. The Wigner function approach systematically involves quantum corrections appearing in the CKT, and definitely keeps the covariance of the fundamental theory Hidaka:2016yjf even in the general coordinate system Liu:2018xip . In this paper, thus we study the semiclassical transport theory with gravitational Riemann curvatures, based on the CKT derived with the Wigner function.

In the following, we present a summary of the findings in this paper. First, we solve the collisionless CKT in general coordinate and derive the Wigner function of Weyl fermions up to O(2)O(\hbar^{2}). This is a contrast to the conventional works, where only the O()O(\hbar) quantum correction is taken into account. It is intriguing that the CKT in curved spacetime is systematically solvable even with the O(2)O(\hbar^{2}) contributions, while the O(2)O(\hbar^{2}) electromagnetic effect is not so tamable; the theory suffers from severe infrared divergence, for which so far no correct prescription is found.

The analysis of the higher-order quantum corrections reveals new aspects of the ambiguity underlying the CKT. It is well-known that due to degrees of freedom in terms of the Lorentz frame, the distribution function of chiral fermions cannot be uniquely determined Chen:2014cla . As a result, a frame vector representing such an ambiguity is inevitably introduced Chen:2015gta . From the Wigner function up to O(2)O(\hbar^{2}), we find that on top of the conventional frame vector, there emerges a different frame vector to define the distribution function. These extra degrees of freedom should be irrelevant to the Wigner function and thus physical quantities, as so is the conventional one. This is one of the guiding principles to identify an equilibrium distribution function. Indeed we find that there is no equilibrium solution for the kinetic equation in general curved spacetime. In other words, in general, the O(2)O(\hbar^{2}) CKT under gravity does not reach a global equilibrium without collisions. However, we elucidate that an equilibrium solution is admitted for several gravitational fields, such as the stationary weak one.

The remaining parts are devoted to the evaluation of physical quantities from the Wigner function that we derived before. For instance, the charge current and energy-momentum tensor are given by the momentum integrals, as follows:

Jμ(x)=2pμ(x,p),Tμν(x)=2pp(μν)(x,p)+O(3)\begin{split}J^{\mu}(x)=2\int_{p}\mathcal{R}^{\mu}(x,p)\,,\quad T^{\mu\nu}(x)=2\int_{p}p^{(\mu}\mathcal{R}^{\nu)}(x,p)+O(\hbar^{3})\end{split} (1)

with p:=d4p(2π)4g(x)\int_{p}:=\int\frac{d^{4}p}{(2\pi)^{4}\sqrt{-g(x)}} and X(μYν)=12(XμYν+YμXν)X^{(\mu}Y^{\nu)}=\frac{1}{2}(X^{\mu}Y^{\nu}+Y^{\mu}X^{\nu}). Here μ(x,p)\mathcal{R}^{\mu}(x,p) is the Wigner function of the right-handed Weyl fermions. Under a static weak gravity, the O()O(\hbar) part of Eq. (1) correctly reproduces the CVE. We find that under a time-dependent gravity, the CVE totally vanishes in the dynamical limit of the background gravitational field. Although this fact is originally derived with the diagrammatic computation Landsteiner:2013aba , our calculation is its first verification based on the CKT.

charge current C0jμenergy-momentum tensor C1tμνstaticjμ=112Rμαξα124ξμR+16ξμRαβξαξβtμν=112Rμν112Rξμξν+124Rημν16Rα(μξν)ξα+16Rαβξαξβ(4ξμξνημν)+16Rμανβξαξβdynamicaljμ=120Rμαξα140ξμRtμν=112Rμν+1105Rξμξν+13840Rημν+115Rα(μξν)ξα2105Rαβξαξβξμξν170Rαβξαξβημν+130Rμανβξαξβ{\begin{array}[]{|\c||\c|\c|}\hline\cr&\text{charge current $C_{0}j^{\mu}$}&\text{energy-momentum tensor $C_{1}t^{\mu\nu}$}\\ \hline\cr\hline\cr\text{static}&{\begin{array}[]{l}j^{\mu}=\frac{1}{12}{R^{\mu}}_{\alpha}\xi^{\alpha}-\frac{1}{24}\xi^{\mu}R\\ \qquad\ +\frac{1}{6}\xi^{\mu}R_{\alpha\beta}\xi^{\alpha}\xi^{\beta}\end{array}}&{\begin{array}[]{l}t^{\mu\nu}=-\frac{1}{12}R^{\mu\nu}-\frac{1}{12}R\xi^{\mu}\xi^{\nu}+\frac{1}{24}R\eta^{\mu\nu}\\ \qquad\quad-\frac{1}{6}R^{\alpha(\mu}\xi^{\nu)}\xi_{\alpha}+\frac{1}{6}R^{\alpha\beta}\xi_{\alpha}\xi_{\beta}(4\xi^{\mu}\xi^{\nu}-\eta^{\mu\nu})\\ \qquad\quad+\frac{1}{6}R^{\mu\alpha\nu\beta}\xi_{\alpha}\xi_{\beta}\end{array}}\\ \hline\cr\text{dynamical}&j^{\mu}=\frac{1}{20}{R^{\mu}}_{\alpha}\xi^{\alpha}-\frac{1}{40}\xi^{\mu}R&{\begin{array}[]{l}t^{\mu\nu}=-\frac{1}{12}R^{\mu\nu}+\frac{1}{105}R\xi^{\mu}\xi^{\nu}+\frac{13}{840}R\eta^{\mu\nu}\\ \qquad\ \ +\frac{1}{15}R^{\alpha(\mu}\xi^{\nu)}\xi_{\alpha}-\frac{2}{105}R^{\alpha\beta}\xi_{\alpha}\xi_{\beta}\xi^{\mu}\xi^{\nu}\\ \qquad\ \ -\frac{1}{70}R^{\alpha\beta}\xi_{\alpha}\xi_{\beta}\eta^{\mu\nu}+\frac{1}{30}R^{\mu\alpha\nu\beta}\xi_{\alpha}\xi_{\beta}\end{array}}\\ \hline\cr\end{array}}
Table 1: Transport phenomena induced by general weak gravity. The static and dynamical parts are derived in Eqs. (78) and (119), respectively.
charge current C0jμenergy-momentum tensor C1tμνstaticj0=2T6T¯,ji=112(×𝝎)it00=2T6T¯,t0i=16(×𝝎)i,tij=112T¯(ij+ηij2)Tdynamicalj0=0,ji=120(×𝝎)it00=0,t0i=120(×𝝎)i,tij=120T¯(ij+ηij2)T+1200σij{\begin{array}[]{|c||c|c|}\hline\cr&\text{charge current $C_{0}j^{\mu}$}&\text{energy-momentum tensor $C_{1}t^{\mu\nu}$}\\ \hline\cr\hline\cr\text{static}&\ j^{0}=\frac{{\boldsymbol{\nabla}}^{2}T}{6\bar{T}}\,,\quad j^{i}=\frac{1}{12}({\boldsymbol{\nabla}}\times{\boldsymbol{\omega}})^{i}&\begin{array}[]{c}t^{00}=\frac{{\boldsymbol{\nabla}}^{2}T}{6\bar{T}}\,,\quad t^{0i}=-\frac{1}{6}({\boldsymbol{\nabla}}\times{\boldsymbol{\omega}})^{i}\,,\\ t^{ij}=-\frac{1}{12\bar{T}}(\partial^{i}\partial^{j}+\eta^{ij}{\boldsymbol{\nabla}}^{2})T\end{array}\\ \hline\cr\text{dynamical}&j^{0}=0\,,\quad j^{i}=\frac{1}{20}({\boldsymbol{\nabla}}\times{\boldsymbol{\omega}})^{i}&\begin{array}[]{c}t^{00}=0\,,\quad t^{0i}=-\frac{1}{20}({\boldsymbol{\nabla}}\times{\boldsymbol{\omega}})^{i}\,,\\ \ t^{ij}=\frac{1}{20\bar{T}}(\partial^{i}\partial^{j}+\eta^{ij}{\boldsymbol{\nabla}}^{2})T+\frac{1}{20}\partial_{0}\sigma^{ij}\end{array}\\ \hline\cr\end{array}}
Table 2: Transport phenomena induced by weak fluid vorticity and temperature gradient. The static and dynamical parts are derived in Eqs. (7) and (7), respectively.

The O(2)O(\hbar^{2}) contribution of Eq. (1) corresponds to the novel transport phenomena induced by Riemann curvatures, which is also our main finding. They are represented as the following charge current and energy-momentum tensor:

Jμ=C1jμ,Tμν=C0tμν.J^{\mu}=C_{1}j^{\mu}\,,\quad T^{\mu\nu}=C_{0}t^{\mu\nu}\,. (2)

Here C0=μ/(2π2)C_{0}=\mu/(2\pi^{2}) and C1=μ2/(4π2)+T2/12C_{1}=\mu^{2}/(4\pi^{2})+T^{2}/12 are the metric-independent coefficients determined by temperature TT and chemical potential for right-handed fermions μ\mu. Also jμj^{\mu} and tμνt^{\mu\nu} are functions of Riemann curvature RμνρσR_{\mu\nu\rho\sigma}, Ricci tensor RμνR_{\mu\nu} and Ricci scalar RR. In Table 2, we summarize the analytic forms of jμj^{\mu} and tμνt^{\mu\nu}. The ‘static’ (‘dynamical’) implies that a time-independent (time-dependent) metric tensor is used. We denote ξμ=δ0μ\xi^{\mu}=\delta^{\mu}_{0} and ημν\eta_{\mu\nu} being the Minkowski metric tensor. It is demonstrated in Appendix D that the same charge current under a static gravity is also derived from different field-theoretical approaches. This consistency apparently guarantees the validity of our formalism in this paper.

The fermionic system under a background fluid is a pedagogical and informative environment to demonstrate the aforementioned novel phenomena. In general the fluid effect is translated into an effective curved spacetime described by the following metric tensor:

g00=1+h00(t,𝒙),g0i=h0i(t,𝒙),gij=ηij,g_{00}=1+h_{00}(t,{\boldsymbol{x}})\,,\quad g_{0i}=h_{0i}(t,{\boldsymbol{x}})\,,\quad g_{ij}=\eta_{ij}\,, (3)

where hμνh_{\mu\nu} is the fluctuation around the flat spacetime. With this metric, temperature gradient and vorticity are given by

iT/T¯=12ih00,ωi=12ε0ijkjhk0.\partial_{i}T/\bar{T}=-\frac{1}{2}\partial_{i}h_{00}\,,\quad\omega^{i}=-\frac{1}{2}\varepsilon^{0ijk}\partial_{j}h_{k0}\,. (4)

Applying Eq. (3) to jμj^{\mu} and tμνt^{\mu\nu} in Table 2, we get the transport phenomena induced by temperature gradient or the inhomogeneity of vorticity. The results are summarized in Table 2. Here we define the shear tensor as σij=ϵij13ηijϵkk\sigma^{ij}=\epsilon^{ij}-\frac{1}{3}\eta^{ij}{\epsilon_{k}}^{k} with ϵij=(ihj)0\epsilon_{ij}=\partial_{(i}h_{j)0}.

There are two crucial features of the novel transport phenomena induced by gravity (or equivalently inhomogeneous fluid profiles). One is that even if the collisionless kinetic equation holds, these transport phenomena are induced, as so are the CME and CVE. Therefore these do not generate any entropy, and thus be nondissipative. This fact motivates us to analyze the relation of Tables 2 and 2 to the quantum anomaly, from different approaches, such as hydrodynamics. This is an interesting open question that will be revisited in the future. The other is the antiparallel flows of the charge current jij^{i} and energy current t0it^{0i}. Tables 2 and 2 for μ>0\mu>0 show the coefficients of these currents have opposite signs whether the metric is static or dynamical and whatever the metric tensor is. This is never explained by the classical particle motions; both charge and energy are carried along the classical particle momenta. The essential ingredient of the antiparallel flow is the spin-Riemann-curvature coupling. To find a more intuitive explanation of such curious flows is also a fascinating task.

In these respects, the novel gravity-induced transport phenomena should involve a lot of implications, e.g., in heavy-ion collisions or Dirac/Weyl semimetals (some of them are discussed in this paper). The latter systems may provide a good playground to study our novel phenomena and can be complementary environments to the former. For them, we need more detailed analysis based on the hydrodynamic model calculation, and quantitative comparison between theory and experiments. Besides, the CKT in curved spacetime and the resulting curvature-induced transport phenomena could play a more crucial role under genuine gravity, although we do not discuss it in this paper. For example, we can discuss the geodesics deviation of chiral fermions due to the spin-curvature coupling. Such a deviation may lead to some correction to the gravitational lensing of neutrinos liebes1964gravitational . Also, the present work could be applicable to the physics of core-collapse supernova explosions and neutron star formations Yamamoto:2015gzz . In this direction, we need to take the collisional effects into account Hidaka:2016yjf ; Hidaka:2017auj ; Yamamoto:2020zrs , based on the Kadanoff-Baym equation in curved spacetime, which respects the diffeomorphism covariance Hohenegger:2008zk .

This paper is organized as follows. In Sec. 2, solving the constraint equations, we obtain the general solution of the Wigner function up to O(2)O(\hbar^{2}) under general curved spacetime. In Sec. 3, we determine an equilibrium distribution function involving O(2)O(\hbar^{2}) corrections, based on the frame-independence of the Wigner function. In Sec. 4, we obtain curvature-induced charge current and energy-momentum tensor in equilibrium, which is consistent with different field-theoretical approaches in Appendix D. In Secs. 5 and 6, we analyze the dynamical response from background gravitational fields. As a practical application, in Sec. 7 we argue the transport phenomena in a fluid with inhomogeneous vorticity and temperature, which yields effective gravitational curvatures. In particular, we observe the antiparallel flows of charge and energy due to an inhomogeneous vorticity (or the Ricci tensor R0i{R_{0}}^{i}). Through this paper, the convention follows from Ref. Liu:2018xip .

2 Chiral kinetic theory at O(2)O(\hbar^{2})

We first show the brief outline of the derivation of the CKT in the Wigner function approach. The kinetic theory is a perturbative effective theory in the infrared momentum regime. In quantum field theory, this perturbation is equivalent to the semiclassical truncation Elze:1986qd . The CKT is obtained from the semiclassical expansions of the equation of motions for the Wigner functions, that is, the Wigner transformed Dyson-Schwinger equation. The Wigner function of chiral fermions obeys three equations, which correspond to Eqs. (8)-(10) in this paper. Two of them are the constraints to the Wigner function, and the other becomes the kinetic equation eventually. In the classical limit, the equation for the Wigner function reduces to the Boltzmann equation (27). The Wigner function formalism gives the quantum generalization of the classical kinetic theory.

Let us start from the Wigner function for the right-handed Weyl fermions, which is defined as Fonarev:1993ht

μ(x,p)=12tr[γμ1+γ52W(x,p)],\displaystyle\displaystyle\mathcal{R}^{\mu}(x,p)=\frac{1}{2}\text{tr}\biggl{[}\gamma^{\mu}\frac{1+\gamma^{5}}{2}W(x,p)\biggr{]}\,, (5)
Wab(x,p)=d4yg(x)eipy/ψ¯b(x,y/2)ψa(x,y/2),\displaystyle\displaystyle W_{ab}(x,p)=\int d^{4}y\,\sqrt{-g(x)}\,e^{-ip\cdot y/\hbar}\langle\bar{\psi}_{b}(x,y/2)\psi_{a}(x,-y/2)\rangle\,,\ \ \ (6)

with g(x)=det(gμν)g(x)=\det(g_{\mu\nu}), ψ¯(x):=ψ(x)γ0^\bar{\psi}(x):=\psi^{\dagger}(x)\gamma^{\hat{0}}, ψ(x,y)=exp(yD)ψ(x)\psi(x,y)=\exp(y\cdot D)\psi(x), ψ¯(x,y)=ψ¯(x)exp(yD)\bar{\psi}(x,y)=\bar{\psi}(x)\exp(y\cdot\overleftarrow{D}), and ψ¯O:=[Oψ]γ0^\bar{\psi}\overleftarrow{O}:=[O\psi]^{\dagger}\gamma^{\hat{0}}. The above W(x,p)W(x,p) is the general relativistic extension of the one in gauge theory Elze:1986qd . Here DμD_{\mu} is called the horizontal lift; for a function on (xμ,yμ)(x^{\mu},y^{\mu}) and (xμ,pμ)(x^{\mu},p_{\mu}), the horizontal lift is represented as

Dμ={μΓμνρyνρy,μ+Γμνρpρpν,D_{\mu}=\begin{cases}\nabla_{\mu}-\Gamma_{\mu\nu}^{\rho}y^{\nu}\partial_{\rho}^{y}\,,\\ \nabla_{\mu}+\Gamma_{\mu\nu}^{\rho}p_{\rho}\partial_{p}^{\nu}\,,\end{cases} (7)

where μ\nabla_{\mu} is the covariant derivative in terms of diffeomorphism and the local Lorentz transformation. The most beneficial property of DμD_{\mu} is that it commutes with both yμy^{\mu} and pμp_{\mu}, while μ\nabla_{\mu} does not.

Hereafter we consider the Dirac theory under an external torsionless gravitational field. In this paper, we focus on the collisionless kinetic theory. The Dirac equation is thus given by γμμψ(x)=0\gamma^{\mu}\nabla_{\mu}\psi(x)=0, which brings the dynamical equation that the Wigner function W(x,p)W(x,p) obeys. After a long computation, the set of equations for μ(x,p)\mathcal{R}^{\mu}(x,p) is up to O(2)O(\hbar^{2}) given by Liu:2018xip

(Dμ+2Pμ)μ=0,\displaystyle(D_{\mu}+\hbar^{2}P_{\mu})\mathcal{R}^{\mu}=0\,, (8)
(pμ+2Qμ)μ=0,\displaystyle(p_{\mu}+\hbar^{2}Q_{\mu})\mathcal{R}^{\mu}=0\,, (9)
εμνρσDρσ+4[(p[μ+2T[μ)ν]+2Sαμνα]=0,\displaystyle\displaystyle\hbar\varepsilon_{\mu\nu\rho\sigma}D^{\rho}\mathcal{R}^{\sigma}+4\Bigl{[}(p_{[\mu}+\hbar^{2}T_{[\mu})\mathcal{R}_{\nu]}+\hbar^{2}S_{\alpha\mu\nu}\mathcal{R}^{\alpha}\Bigr{]}=0\,, (10)

where we introduce the following notations:

Pμ=18λRμνpλpν124λRρσμνpλpνpσpρ+18RρσμνpνpσDρ,\displaystyle\displaystyle P_{\mu}=-\frac{1}{8}\nabla_{\lambda}R_{\mu\nu}\partial_{p}^{\lambda}\partial_{p}^{\nu}-\frac{1}{24}\nabla_{\lambda}{R^{\rho}}_{\sigma\mu\nu}\partial_{p}^{\lambda}\partial_{p}^{\nu}\partial_{p}^{\sigma}p_{\rho}+\frac{1}{8}{R^{\rho}}_{\sigma\mu\nu}\partial_{p}^{\nu}\partial_{p}^{\sigma}D_{\rho}\,, (11)
Qμ=18Rμνpν+124Rρσμνpνpσpρ=3Aμ+Bμ,\displaystyle\displaystyle Q_{\mu}=\frac{1}{8}R_{\mu\nu}\partial_{p}^{\nu}+\frac{1}{24}{R^{\rho}}_{\sigma\mu\nu}\partial^{\nu}_{p}\partial_{p}^{\sigma}p_{\rho}=3A_{\mu}+B_{\mu}\,, (12)
Tμ=14Rμνpν+124Rρσμνpνpσpρ=6Aμ+Bμ,\displaystyle\displaystyle T_{\mu}=\frac{1}{4}R_{\mu\nu}\partial_{p}^{\nu}+\frac{1}{24}{R^{\rho}}_{\sigma\mu\nu}\partial^{\nu}_{p}\partial_{p}^{\sigma}p_{\rho}=6A_{\mu}+B_{\mu}\,, (13)
Aμ=124Rμνpν,Bμ=124Rρσμνpνpσpρ,Sαμν=116Rλαμνpλ.\displaystyle\displaystyle A_{\mu}=\frac{1}{24}R_{\mu\nu}\partial_{p}^{\nu}\,,\quad B_{\mu}=\frac{1}{24}{R^{\rho}}_{\sigma\mu\nu}\partial_{p}^{\nu}\partial_{p}^{\sigma}p_{\rho}\,,\quad S_{\alpha\mu\nu}=-\frac{1}{16}R_{\lambda\alpha\mu\nu}\partial^{\lambda}_{p}\,. (14)

In the above equations, we denote X[μYν]=(XμYνXνYμ)/2X^{[\mu}Y^{\nu]}=(X^{\mu}Y^{\nu}-X^{\nu}Y^{\mu})/2, the Riemann tensor is defined as Rρσμν=2([νΓμ]σρ+Γλ[νρΓμ]σλ){R^{\rho}}_{\sigma\mu\nu}=2\bigl{(}\partial_{[\nu}\Gamma^{\rho}_{\mu]\sigma}+\Gamma_{\lambda[\nu}^{\rho}\Gamma_{\mu]\sigma}^{\lambda}\bigr{)} with Γμνρ=gρλ(μgλν+νgλμλgμν)/2\Gamma^{\rho}_{\mu\nu}=g^{\rho\lambda}(\partial_{\mu}g_{\lambda\nu}+\partial_{\nu}g_{\lambda\mu}-\partial_{\lambda}g_{\mu\nu})/2, and the Ricci tensor is Rμν=RλμλνR_{\mu\nu}={R^{\lambda}}_{\mu\lambda\nu}. For left-handed Weyl fermions, similar equations are derived, but only the sign in front of εμνρσ\varepsilon_{\mu\nu\rho\sigma} is flipped, as is parity-odd. The first equation (8) corresponds to the kinetic equation while the others (9) and (10) are constraints that determine the functional form of μ\mathcal{R}^{\mu}. It is worthwhile to mention that Eqs. (8)-(10) are the Ward identities in terms of the symmetries that Weyl fermions respect in a given coordinate; the U(1)U(1) gauge symmetry, the conformal symmetry, and the Lorentz symmetry, respectively Liu:2020flb .

Let us parametrize the solution for Eqs. (9) and (10) as

μ=(0)μ+(1)μ+2(2)μ.\mathcal{R}^{\mu}=\mathcal{R}^{\mu}_{(0)}+\hbar\mathcal{R}^{\mu}_{(1)}+\hbar^{2}\mathcal{R}_{(2)}^{\mu}\,. (15)

Contracting Eq. (10) with pνp^{\nu}, we find

p2μ=pμp+2εμνρσpνDρσ+22pν(T[μν]+Sαμνα).p^{2}\mathcal{R}_{\mu}=p_{\mu}p\cdot\mathcal{R}+\frac{\hbar}{2}\varepsilon_{\mu\nu\rho\sigma}p^{\nu}D^{\rho}\mathcal{R}^{\sigma}+2\hbar^{2}p^{\nu}\Bigl{(}T_{[\mu}\mathcal{R}_{\nu]}+S_{\alpha\mu\nu}\mathcal{R}^{\alpha}\Bigr{)}\,. (16)

Combined with Eq. (9), this equation is decomposed into

p2μ(0)=0,\displaystyle p^{2}\mathcal{R}_{\mu}^{(0)}=0\,, (17)
p2μ(1)=12εμνρσpνDρ(0)σ,\displaystyle\displaystyle p^{2}\mathcal{R}_{\mu}^{(1)}=\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}p^{\nu}D^{\rho}\mathcal{R}^{\sigma}_{(0)}\,, (18)
p2μ(2)=pμQ(0)+12εμνρσpνDρ(1)σ+2pν(T[μν](0)+Sαμν(0)α).\displaystyle\displaystyle p^{2}\mathcal{R}_{\mu}^{(2)}=-p_{\mu}Q\cdot\mathcal{R}_{(0)}+\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}p^{\nu}D^{\rho}\mathcal{R}^{\sigma}_{(1)}+2p^{\nu}\Bigl{(}T_{[\mu}\mathcal{R}_{\nu]}^{(0)}+S_{\alpha\mu\nu}\mathcal{R}^{\alpha}_{(0)}\Bigr{)}\,. (19)

Also Eqs. (9) and (10) yield

p(0)=0,\displaystyle p\cdot\mathcal{R}_{(0)}=0\,, (20)
p(1)=0,\displaystyle p\cdot\mathcal{R}_{(1)}=0\,, (21)
p(2)+Q(0)=0,\displaystyle p\cdot\mathcal{R}_{(2)}+Q\cdot\mathcal{R}_{(0)}=0\,, (22)
4p[μν](0)=0,\displaystyle 4p_{[\mu}\mathcal{R}^{(0)}_{\nu]}=0\,, (23)
4p[μν](1)+εμνρσDρ(0)σ=0,\displaystyle 4p_{[\mu}\mathcal{R}^{(1)}_{\nu]}+\varepsilon_{\mu\nu\rho\sigma}D^{\rho}\mathcal{R}^{\sigma}_{(0)}=0\,, (24)
4p[μν](2)+εμνρσDρ(1)σ+4(T[μν](0)+Sαμν(0)α)=0.\displaystyle 4p_{[\mu}\mathcal{R}_{\nu]}^{(2)}+\varepsilon_{\mu\nu\rho\sigma}D^{\rho}\mathcal{R}_{(1)}^{\sigma}+4\Bigl{(}T_{[\mu}\mathcal{R}_{\nu]}^{(0)}+S_{\alpha\mu\nu}\mathcal{R}_{(0)}^{\alpha}\Bigr{)}=0\,. (25)

In the following, we look for (0)μ\mathcal{R}^{\mu}_{(0)}, (1)μ\mathcal{R}^{\mu}_{(1)} and (2)μ\mathcal{R}^{\mu}_{(2)} that satisfy Eqs. (17)-(25).

First, let us solve the zeroth and first-order parts. Equations (17) and (20) imply

(0)μ=2πδ(p2)pμf(0),\mathcal{R}^{\mu}_{(0)}=2\pi\delta(p^{2})p^{\mu}f_{(0)}\,, (26)

where f(0)f_{(0)} is a scalar function that satisfies δ(p2)p2f(0)=0\delta(p^{2})p^{2}f_{(0)}=0. From Eq. (23), we can check that there does not appear any other term in μ(0)\mathcal{R}_{\mu}^{(0)}. Equation (8) in the zeroth order, Dμ(0)μ=0D_{\mu}\mathcal{R}^{\mu}_{(0)}=0, gives the collisionless Boltzmann equation,

2πδ(p2)(pμμ+Γμνρpμpρpν)f(0)=0.2\pi\delta(p^{2})(p^{\mu}\nabla_{\mu}+\Gamma_{\mu\nu}^{\rho}p^{\mu}p_{\rho}\partial_{p}^{\nu})f_{(0)}=0\,. (27)

Higher-order terms give quantum corrections to the Boltzmann equation.

From Eq. (18) and the above (0)μ\mathcal{R}^{\mu}_{(0)}, we find

p2(1)μ=0.p^{2}\mathcal{R}^{\mu}_{(1)}=0\,. (28)

This does not necessarily mean that (1)μ\mathcal{R}^{\mu}_{(1)} itself vanishes for arbitrary pμp_{\mu}. Indeed if (1)μ\mathcal{R}^{\mu}_{(1)} involves δ(p2)\delta(p^{2}), it fulfils Eq. (28). Therefore, the first-order correction is generally written as

(1)μ=2πδ(p2)~(1)μ.\mathcal{R}^{\mu}_{(1)}=2\pi\delta(p^{2})\widetilde{\mathcal{R}}^{\mu}_{(1)}\,. (29)

Here the undetermined part ~μ(1)\widetilde{\mathcal{R}}_{\mu}^{(1)} satisfies δ(p2)p2~(1)μ=0\delta(p^{2})p^{2}\widetilde{\mathcal{R}}^{\mu}_{(1)}=0 so that Eq. (28) holds. Plugging this (1)μ\mathcal{R}^{\mu}_{(1)} and (0)μ\mathcal{R}^{\mu}_{(0)} into Eq. (24), we obtain

δ(p2)[εμνρσpρDσf(0)4p[μ~ν](1)]=0.\delta(p^{2})\Bigl{[}\varepsilon_{\mu\nu\rho\sigma}p^{\rho}D^{\sigma}f_{(0)}-4p_{[\mu}\widetilde{\mathcal{R}}_{\nu]}^{(1)}\Bigr{]}=0\,. (30)

We contract this with nν/(2pn)n^{\nu}/(2\,p\cdot n), where nμ(x)n^{\mu}(x) is a vector field independent of pμp_{\mu}. Then we get

~μ(1)δ(p2)=δ(p2)[pμn~(1)pn+εμνρσpρnσ2pnDνf(0)].\begin{split}\widetilde{\mathcal{R}}_{\mu}^{(1)}\delta(p^{2})=\delta(p^{2})\biggl{[}p_{\mu}\frac{n\cdot\widetilde{\mathcal{R}}_{(1)}}{p\cdot n}+\frac{\varepsilon_{\mu\nu\rho\sigma}p^{\rho}n^{\sigma}}{2p\cdot n}D^{\nu}f_{(0)}\biggr{]}\,.\end{split} (31)

Thus the first-order correction is given by

(1)μ=2πδ(p2)[pμf(1)+ΣnμνDνf(0)],\mathcal{R}^{\mu}_{(1)}=2\pi\delta(p^{2})\Bigl{[}p^{\mu}f_{(1)}+\Sigma_{n}^{\mu\nu}D_{\nu}f_{(0)}\Bigr{]}\,, (32)

where we define

f(1)=n~(1)pn,Σnμν=εμνρσpρnσ2pn.f_{(1)}=\frac{n\cdot\widetilde{\mathcal{R}}^{(1)}}{p\cdot n}\,,\quad\Sigma_{n}^{\mu\nu}=\frac{\varepsilon^{\mu\nu\rho\sigma}p_{\rho}n_{\sigma}}{2p\cdot n}\,. (33)

In the above (1)μ\mathcal{R}^{\mu}_{{(1)}}, an arbitrary vector nμn^{\mu} emerges through Σnμν\Sigma^{\mu\nu}_{n}. This ambiguity is related to the (local) Lorentz transformation Chen:2014cla , and thus Σnμν\Sigma^{\mu\nu}_{n} is regarded as the spin tensor defined in the frame nμn^{\mu} Chen:2015gta . In particular, at nμ=(1,𝟎)n^{\mu}=(1,{\boldsymbol{0}}), we have f(1)=~0(1)/p0f_{(1)}=\widetilde{\mathcal{R}}^{(1)}_{0}/p_{0}, i.e., the charge density divided by the particle energy. In this sense, f(1)f_{(1)} can be regarded as the quantum correction to the distribution function. Note that the solution (1)μ\mathcal{R}^{\mu}_{(1)} fulfils Eqs. (18) and (21) as long as δ(p2)p2f(1)=0\delta(p^{2})p^{2}f_{(1)}=0 holds.

Now we solve the second-order correction. By plugging the above (0)μ\mathcal{R}_{(0)}^{\mu} and (1)μ\mathcal{R}_{(1)}^{\mu} into Eq. (19), we obtain

p2μ(2)=2π(pμQp+pν𝒟μν)δ(p2)f(0),\begin{split}p^{2}\mathcal{R}_{\mu}^{(2)}&=2\pi\Bigl{(}-p_{\mu}Q\cdot p+p^{\nu}\mathcal{D}_{\mu\nu}\Bigr{)}\delta(p^{2})f_{(0)}\,,\end{split} (34)

where the derivative operator 𝒟μν\mathcal{D}_{\mu\nu} is defined as

𝒟μν=2(T[μpν]+Sαμνpα)+12εμνρσDρΣnσλDλ.\mathcal{D}_{\mu\nu}=2\Bigl{(}T_{[\mu}p_{\nu]}+S_{\alpha\mu\nu}p^{\alpha}\Bigr{)}+\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}D^{\rho}\Sigma_{n}^{\sigma\lambda}D_{\lambda}\,. (35)

The general form of the second-order correction then reads

μ(2)=2πδ(p2)~μ(2)+2πp2[pμQp+pν𝒟μν]δ(p2)f(0).\mathcal{R}_{\mu}^{(2)}=2\pi\delta(p^{2})\widetilde{\mathcal{R}}_{\mu}^{(2)}+\frac{2\pi}{p^{2}}\biggl{[}-p_{\mu}Q\cdot p+p^{\nu}\mathcal{D}_{\mu\nu}\biggr{]}\delta(p^{2})f_{(0)}\,. (36)

Here we again introduced the undetermined part ~μ(2)\widetilde{\mathcal{R}}_{\mu}^{(2)}, which satisfies δ(p2)p2~(2)μ=0\delta(p^{2})p^{2}\widetilde{\mathcal{R}}^{\mu}_{(2)}=0. Plugging μ(0)\mathcal{R}_{\mu}^{(0)}, μ(1)\mathcal{R}_{\mu}^{(1)}, and μ(2)\mathcal{R}_{\mu}^{(2)} into Eq. (25), we obtain

0=4p[μν](2)+(2π)εμνρσDρ(pσf(1)+ΣnσλDλf(0))δ(p2)+4(2π)(T[μpν]+Sαμνpα)f(0)δ(p2)=2π[4p[μ~ν](2)+εμνρσDρpσf(1)+4p2p[μpρ𝒟ν]ρf(0)+2𝒟μνf(0)]δ(p2)=2π[4p[μ~ν](2)εμνρσpρDσf(1)1p2εμνρσpρεαβγσpα𝒟βγf(0)]δ(p2).\begin{split}0&=4p_{[\mu}\mathcal{R}^{(2)}_{\nu]}+(2\pi)\varepsilon_{\mu\nu\rho\sigma}D^{\rho}\Bigl{(}p^{\sigma}f_{(1)}+\Sigma_{n}^{\sigma\lambda}D_{\lambda}f_{(0)}\Bigr{)}\delta(p^{2})+4(2\pi)\Bigl{(}T_{[\mu}p_{\nu]}+S_{\alpha\mu\nu}p^{\alpha}\Bigr{)}f_{(0)}\delta(p^{2})\\ &=2\pi\biggl{[}4p_{[\mu}\widetilde{\mathcal{R}}^{(2)}_{\nu]}+\varepsilon_{\mu\nu\rho\sigma}D^{\rho}p^{\sigma}f_{(1)}+\frac{4}{p^{2}}p_{[\mu}p^{\rho}\mathcal{D}_{\nu]\rho}f_{(0)}+2\mathcal{D}_{\mu\nu}f_{(0)}\Bigr{]}\delta(p^{2})\\ &=2\pi\biggl{[}4p_{[\mu}\widetilde{\mathcal{R}}^{(2)}_{\nu]}-\varepsilon_{\mu\nu\rho\sigma}p^{\rho}D^{\sigma}f_{(1)}-\frac{1}{p^{2}}\varepsilon_{\mu\nu\rho\sigma}p^{\rho}\varepsilon^{\alpha\beta\gamma\sigma}p_{\alpha}\mathcal{D}_{\beta\gamma}f_{(0)}\biggr{]}\delta(p^{2})\,.\end{split} (37)

Similarly to Eq. (30), we solve the above equation by introducing a vector uμu^{\mu} (this is in general different from nμn^{\mu}), as follows:

~μ(2)δ(p2)=δ(p2)[pμf(2)+ΣμνuDνf(1)]+1p2εαβγνΣμνupα𝒟βγδ(p2)f(0)=δ(p2)[pμf(2)+ΣμνuDνf(1)]δ(p2)p2Σμνu[12R~αβνρpρpαpβ+pDΣnνρDρ]f(0),\begin{split}\widetilde{\mathcal{R}}_{\mu}^{(2)}\delta(p^{2})&=\delta(p^{2})\Bigl{[}p_{\mu}f_{(2)}+\Sigma_{\mu\nu}^{u}D^{\nu}f_{(1)}\Bigr{]}+\frac{1}{p^{2}}\varepsilon^{\alpha\beta\gamma\nu}\Sigma_{\mu\nu}^{u}p_{\alpha}\mathcal{D}_{\beta\gamma}\delta(p^{2})f_{(0)}\\ &=\delta(p^{2})\Bigl{[}p_{\mu}f_{(2)}+\Sigma_{\mu\nu}^{u}D^{\nu}f_{(1)}\Bigr{]}-\frac{\delta(p^{2})}{p^{2}}\Sigma_{\mu\nu}^{u}\biggl{[}\frac{1}{2}\tilde{R}^{\alpha\beta\nu\rho}p^{\rho}p^{\alpha}\partial_{p}^{\beta}+p\cdot D\Sigma_{n}^{\nu\rho}D_{\rho}\biggr{]}f_{(0)}\,,\end{split} (38)

where we defined R~αβμν=Rαβρσερσμν/2\tilde{R}_{\alpha\beta\mu\nu}={R_{\alpha\beta}}^{\rho\sigma}\varepsilon_{\rho\sigma\mu\nu}/2 and

f(2)=u~(2)pu.f_{(2)}=\frac{u\cdot\widetilde{\mathcal{R}}_{(2)}}{p\cdot u}\,. (39)

In the second line of Eq. (38), we utilized

[Aμ,pν]=124Rμν,[Bμ,pν]=124Rμν+124(Rρνμσ+Rρσμν)pρpσ,\begin{split}[A_{\mu},p_{\nu}]=\frac{1}{24}R_{\mu\nu}\,,\qquad[B_{\mu},p_{\nu}]=-\frac{1}{24}R_{\mu\nu}+\frac{1}{24}\Bigl{(}{R^{\rho}}_{\nu\mu\sigma}+{R^{\rho}}_{\sigma\mu\nu}\Bigr{)}p_{\rho}\partial_{p}^{\sigma}\,,\end{split} (40)

which yield

2εαβγνpα(Tβpγ+Sλβγpλ)=12R~αβνρpρpαβp.\begin{split}2\varepsilon^{\alpha\beta\gamma\nu}p_{\alpha}\Bigl{(}T_{\beta}p_{\gamma}+S_{\lambda\beta\gamma}p^{\lambda}\Bigr{)}&=-\frac{1}{2}\tilde{R}^{\alpha\beta\nu\rho}p_{\rho}p_{\alpha}\partial_{\beta}^{p}\,.\end{split} (41)

Therefore, the second-order correction reads

μ(2)=2πδ(p2)[pμf(2)+ΣμνuDνf(1)]+2π1p2[pμQp+2pν(T[μpν]+Sαμνpα)]δ(p2)f(0)+2πδ(p2)p2[12εμνρσpνDρΣnσλDλΣμνu(12R~αβνρpρpαβp+pDΣnνρDρ)]f(0).\begin{split}\mathcal{R}_{\mu}^{(2)}&=2\pi\delta(p^{2})\biggl{[}p_{\mu}f_{(2)}+\Sigma_{\mu\nu}^{u}D^{\nu}f_{(1)}\biggr{]}+2\pi\frac{1}{p^{2}}\biggl{[}-p_{\mu}Q\cdot p+2p^{\nu}\Bigl{(}T_{[\mu}p_{\nu]}+S_{\alpha\mu\nu}p^{\alpha}\Bigr{)}\biggr{]}\delta(p^{2})f_{(0)}\\ &\quad+2\pi\frac{\delta(p^{2})}{p^{2}}\biggl{[}\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}p^{\nu}D^{\rho}\Sigma_{n}^{\sigma\lambda}D_{\lambda}-\Sigma_{\mu\nu}^{u}\biggl{(}\frac{1}{2}{\tilde{R}}^{\alpha\beta\nu\rho}p_{\rho}p_{\alpha}\partial_{\beta}^{p}+p\cdot D\Sigma_{n}^{\nu\rho}D_{\rho}\biggr{)}\biggr{]}f_{(0)}\,.\end{split} (42)

We mention that Eq. (19) is still fulfilled for the above μ(2)\mathcal{R}_{\mu}^{(2)} as long as

δ(p2)p2f(2)=0\delta(p^{2})p^{2}f_{(2)}=0 (43)

holds. Indeed we can check

δ(p2)p2~μ(2)=δ(p2)Σμνu[12R~αβνρpρpαβp+pDΣnνρDρ]f(0)=δ(p2)Σμνu[12R~αβνρpρpαβp+Dλ(Σnνλpρ+12ενλρσpσ12ενλρσp2nσpn)Dρ]f(0)=δ(p2)ΣμνuDλΣnνλpDf(0)=0.\begin{split}\delta(p^{2})p^{2}\widetilde{\mathcal{R}}^{(2)}_{\mu}&=-\delta(p^{2})\Sigma_{\mu\nu}^{u}\biggl{[}\frac{1}{2}\tilde{R}^{\alpha\beta\nu\rho}p_{\rho}p_{\alpha}\partial^{p}_{\beta}+p\cdot D\Sigma_{n}^{\nu\rho}D_{\rho}\biggr{]}f_{(0)}\\ &=-\delta(p^{2})\Sigma_{\mu\nu}^{u}\biggl{[}\frac{1}{2}\tilde{R}^{\alpha\beta\nu\rho}p_{\rho}p_{\alpha}\partial^{p}_{\beta}+D_{\lambda}\biggl{(}\Sigma_{n}^{\nu\lambda}p^{\rho}+\frac{1}{2}\varepsilon^{\nu\lambda\rho\sigma}p_{\sigma}-\frac{1}{2}\varepsilon^{\nu\lambda\rho\sigma}\frac{p^{2}n_{\sigma}}{p\cdot n}\biggr{)}D_{\rho}\biggr{]}f_{(0)}\\ &=-\delta(p^{2})\Sigma_{\mu\nu}^{u}D_{\lambda}\Sigma_{n}^{\nu\lambda}p\cdot Df_{(0)}=0\,.\end{split} (44)

In the second line, we utilized

Σα[μnpν]=12Σμνnpα14εμναβ(pβp2nβpn),\Sigma_{\alpha[\mu}^{n}p_{\nu]}=-\frac{1}{2}\Sigma_{\mu\nu}^{n}p_{\alpha}-\frac{1}{4}\varepsilon_{\mu\nu\alpha\beta}\biggl{(}p^{\beta}-\frac{p^{2}n^{\beta}}{p\cdot n}\biggr{)}\,, (45)

which follows from the Schouten identity: pμενρσλ+pνερσλμ+pρεσλμν+pσελμνρ+pλεμνρσ=0p_{\mu}\varepsilon_{\nu\rho\sigma\lambda}+p_{\nu}\varepsilon_{\rho\sigma\lambda\mu}+p_{\rho}\varepsilon_{\sigma\lambda\mu\nu}+p_{\sigma}\varepsilon_{\lambda\mu\nu\rho}+p_{\lambda}\varepsilon_{\mu\nu\rho\sigma}=0. Also the last line follows from [Dμ,Dν]f=Rαβμνpαpβf[D_{\mu},D_{\nu}]f=-R_{\alpha\beta\mu\nu}p^{\alpha}\partial^{\beta}_{p}f, and the classical kinetic equation (8), i.e., δ(p2)pDf(0)=0\delta(p^{2})\,p\cdot Df_{(0)}=0. We stress that Eq. (43) is a crucial constraint to f(2)f_{(2)}, especially when we determine the equilibrium distribution function (see Sec. 3).

Eventually, the Wigner function up to O(2)O(\hbar^{2}) is derived as

μ=2πδ(p2)[pμ(f(0)+f(1)+2f(2))+ΣμνnDνf(0)+2ΣμνuDνf(1)]+2π21p2[pμQp+2pν(T[μpν]+Sαμνpα)]δ(p2)f(0)+2π2δ(p2)p2[12εμνρσpνDρΣnσλDλΣμνu(12R~αβνρpρpαβp+pDΣnνρDρ)]f(0).\begin{split}\mathcal{R}_{\mu}&=2\pi\delta(p^{2})\Bigl{[}p_{\mu}\Bigl{(}f_{(0)}+\hbar f_{(1)}+\hbar^{2}f_{(2)}\Bigr{)}+\hbar\Sigma_{\mu\nu}^{n}D^{\nu}f_{(0)}+\hbar^{2}\Sigma_{\mu\nu}^{u}D^{\nu}f_{(1)}\Bigr{]}\\ &\quad+2\pi\hbar^{2}\frac{1}{p^{2}}\biggl{[}-p_{\mu}Q\cdot p+2p^{\nu}\Bigl{(}T_{[\mu}p_{\nu]}+S_{\alpha\mu\nu}p^{\alpha}\Bigr{)}\biggr{]}\delta(p^{2})f_{(0)}\\ &\quad+2\pi\hbar^{2}\frac{\delta(p^{2})}{p^{2}}\biggl{[}\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}p^{\nu}D^{\rho}\Sigma_{n}^{\sigma\lambda}D_{\lambda}-\Sigma_{\mu\nu}^{u}\biggl{(}\frac{1}{2}{\tilde{R}}^{\alpha\beta\nu\rho}p_{\rho}p_{\alpha}\partial_{\beta}^{p}+p\cdot D\Sigma_{n}^{\nu\rho}D_{\rho}\biggr{)}\biggr{]}f_{(0)}\,.\end{split} (46)

As the above μ\mathcal{R}^{\mu} is a fermion propagator, performing the momentum integration involving it we evaluate a corresponding quantity of Weyl fermions under a gravitational field. In particular, the charge current and the symmetric energy-momentum tensor are given by

Jμ(x)=ptr[γμ1+γ52W(x,p)]=2pμ(x,p),Tμν(x)=ptr[i2γ(μDν)1+γ52W(x,p)]=2pp(μν)(x,p)+O(3),\begin{split}J^{\mu}(x)&=\int_{p}\text{tr}\biggl{[}\gamma^{\mu}\frac{1+\gamma^{5}}{2}W(x,p)\biggr{]}=2\int_{p}\mathcal{R}^{\mu}(x,p)\,,\\ \quad T^{\mu\nu}(x)&=\int_{p}\text{tr}\biggl{[}\frac{i\hbar}{2}\gamma^{(\mu}\overleftrightarrow{D}^{\nu)}\frac{1+\gamma^{5}}{2}W(x,p)\biggr{]}=2\int_{p}p^{(\mu}\mathcal{R}^{\nu)}(x,p)+O(\hbar^{3})\,,\end{split} (47)

where we define p:=d4p(2π)4g(x)\int_{p}:=\int\frac{d^{4}p}{(2\pi)^{4}\sqrt{-g(x)}}, X(μYν)=(XμYν+XνYμ)/2X^{(\mu}Y^{\nu)}=(X^{\mu}Y^{\nu}+X^{\nu}Y^{\mu})/2 and Dμ(ψ¯bψa)=ψ¯bDμψaψ¯bDμψa\overleftrightarrow{D_{\mu}}(\bar{\psi}_{b}\psi_{a})=\bar{\psi}_{b}D_{\mu}\psi_{a}-\bar{\psi}_{b}\overleftarrow{D}_{\mu}\psi_{a}. Here we used the expansion of the derivative; Dμψ(x,y)=μyψ(x,y)+O(2)D_{\mu}\psi(x,y)=\partial_{\mu}^{y}\psi(x,y)+O(\hbar^{2}) (see Appendix C in Ref. Liu:2018xip ). In Sec. 4, we derive the equilibrium Wigner function and show that it yields the gravity-induced parts of JμJ^{\mu} and TμνT^{\mu\nu}. We also demonstrate that the different approaches in Appendix D lead to the same JμJ^{\mu} [see Eq. (78) and Eqs. (169) and (186)].

3 Frame dependence and equilibrium

In the evaluation of physical quantities such as Eq. (47), it is necessary to identify the explicit form of f(0),(1),(2)f_{{(0)},{(1)},{(2)}}. For this purpose, the frame (i.e., nμn^{\mu} and uμu^{\mu}) dependences of μ\mathcal{R}^{\mu} are a crucial concept; as shown below, we derive a constraint on the distribution function from the proper transformation law under the shift of the frames. Combined this with the kinetic equation, we determine f(0),(1),(2)f_{{(0)},{(1)},{(2)}} at equilibrium, which are utilized to evaluate equilibrium transport phenomena induced by external gravity in Sec. 4. This section is devoted to the analysis of the frame dependence and the equilibrium solution found from it.

In the above derivation of μ\mathcal{R}^{\mu}, the frame vectors nμn^{\mu} and uμu^{\mu} are algebraically introduced. It is valid to expect that the frame-dependence disappears in μ\mathcal{R}^{\mu}, which generates physical quantities. Indeed, as is well-known in the O()O(\hbar) CKT, the choice of the frame vector nμn^{\mu} corresponds to the Lorentz transformation, and the frame-dependence is totally compensated in physical quantities, due to the shift of f(1)f_{(1)}. Hence we may plausibly require that the same is true in the O(2)O(\hbar^{2}) CKT. That is, we determine the transformation law of f(2)f_{(2)} under nμnμn^{\mu}\to n^{\prime\mu} and uμuμu^{\mu}\to u^{\prime\mu} so that the frame dependence vanishes in μ\mathcal{R}^{\mu}.

Let us first take the Lorentz transformation in terms of nμn^{\mu}, namely, (xμ,pμ)(xμ,pμ)=(Λn)μν(xν,pν)(x^{\mu},p^{\mu})\to(x^{\prime\mu},p^{\prime\mu})={({\Lambda}_{n})^{\mu}}_{\nu}(x^{\nu},p^{\nu}) and uμuμ=(Λn)μνuνu^{\mu}\to u^{\prime\mu}={({\Lambda}_{n})^{\mu}}_{\nu}u^{\nu}, where (Λn)μν{({\Lambda}_{n})^{\mu}}_{\nu} is the matrix representation of the local Lorentz transformation. This transformation is equivalent to the one of the frame vector nμn^{\mu} as

nμnμ=(Λn1)μνnν.n^{\mu}\to n^{\prime\mu}={{({\Lambda}_{n}^{-1})}^{\mu}}_{\nu}n^{\nu}\,. (48)

We also parametrize the transformation of ff as

f(x,p)f(x,p)=f(x,p)+δnf(1)(x,p)+2δnf(2)(x,p).f(x,p)\to f^{\prime}(x^{\prime},p^{\prime})=f(x,p)+\hbar\delta_{n}f_{(1)}(x,p)+\hbar^{2}\delta_{n}f_{(2)}(x,p)\,. (49)

Due to the Lorentz covariance of μ\mathcal{R}^{\mu}, we have

0=(Λn1)μνν(x,p)μ(x,p)=2πδ(p2)[pμ(δnf(1)+2δnf(2))+(ΣμνnΣμνn)Dνf(0)+2ΣμνuDνδnf(1)+2p2(12εμνρσpνDρ(ΣnσλΣnσλ)Dλf(0)ΣμνupD(ΣnνρΣnνρ)Dρf(0))].\begin{split}0&={(\Lambda^{-1}_{n})_{\mu}}^{\nu}\mathcal{R}^{\prime}_{\nu}(x^{\prime},p^{\prime})-\mathcal{R}_{\mu}(x,p)\\ &=2\pi\delta(p^{2})\biggl{[}p_{\mu}\Bigl{(}\hbar\delta_{n}f_{(1)}+\hbar^{2}\delta_{n}f_{(2)}\Bigr{)}+\hbar\Bigl{(}\Sigma^{n^{\prime}}_{\mu\nu}-\Sigma^{n}_{\mu\nu}\Bigr{)}D^{\nu}f_{(0)}+\hbar^{2}\Sigma^{u}_{\mu\nu}D^{\nu}\delta_{n}f_{(1)}\\ &\qquad\qquad\qquad+\frac{\hbar^{2}}{p^{2}}\biggl{(}\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}p^{\nu}D^{\rho}\Bigl{(}\Sigma^{\sigma\lambda}_{n^{\prime}}-\Sigma^{\sigma\lambda}_{n}\Bigr{)}D_{\lambda}f_{(0)}-\Sigma_{\mu\nu}^{u}p\cdot D\Bigl{(}\Sigma_{n^{\prime}}^{\nu\rho}-\Sigma_{n}^{\nu\rho}\Bigr{)}D_{\rho}f_{(0)}\biggr{)}\biggr{]}\,.\end{split} (50)

Contracting Eq. (50) with nμn^{\mu} and picking up only the O()O(\hbar) terms, we find

δnf(1)=nμpnΣμνnDνf(0).\begin{split}\delta_{n}f_{(1)}&=-\frac{n^{\mu}}{p\cdot n}\Sigma_{\mu\nu}^{n^{\prime}}D^{\nu}f_{(0)}\,.\end{split} (51)

Similarly, contracting Eq. (50) with uμu^{\mu}, we obtain

δnf(2)=1p2ΣμνuDμ(ΣnνρΣnνρ)Dρf(0).\begin{split}\delta_{n}f_{(2)}&=\frac{1}{p^{2}}\Sigma_{\mu\nu}^{u}D^{\mu}\Bigl{(}\Sigma^{\nu\rho}_{n^{\prime}}-\Sigma^{\nu\rho}_{n}\Bigr{)}D_{\rho}f_{(0)}\,.\end{split} (52)

The above δnf(1),(2)\delta_{n}f_{{(1)},{(2)}} fulfills δ(p2)p2δnf(1),(2)=0\delta(p^{2})p^{2}\delta_{n}f_{{(1)},{(2)}}=0. Also, we can show that they satisfy Eq. (50).

Let us also perform the Lorentz transformation with

uμuμ=(Λu1)μνuν,u^{\mu}\to u^{\prime\mu}={{({\Lambda}_{u}^{-1})}^{\mu}}_{\nu}u^{\nu}\,, (53)

for which the Lorentz covariance of μ\mathcal{R}_{\mu} requires

0=(Λu1)μνν(x,p)μ(x,p)=2πδ(p2)[pμ(δuf(1)+2δuf(2))+2(ΣμνuΣμνu)Dνf(1)+2ΣμνuDνδuf(1)2p2(ΣμνuΣμνu)(12R~αβνρpρpαβp+pDΣnνρDρ)f(0)].\begin{split}0&={(\Lambda^{-1}_{u})_{\mu}}^{\nu}\mathcal{R}^{\prime}_{\nu}(x^{\prime},p^{\prime})-\mathcal{R}_{\mu}(x,p)\\ &=2\pi\delta(p^{2})\biggl{[}p_{\mu}\Bigl{(}\hbar\delta_{u}f_{(1)}+\hbar^{2}\delta_{u}f_{(2)}\Bigr{)}+\hbar^{2}\Bigl{(}\Sigma^{u^{\prime}}_{\mu\nu}-\Sigma^{u}_{\mu\nu}\Bigr{)}D^{\nu}f_{(1)}+\hbar^{2}\Sigma^{u^{\prime}}_{\mu\nu}D^{\nu}\delta_{u}f_{(1)}\\ &\qquad\qquad\qquad-\frac{\hbar^{2}}{p^{2}}\Bigl{(}\Sigma^{u^{\prime}}_{\mu\nu}-\Sigma^{u}_{\mu\nu}\Bigr{)}\biggl{(}\frac{1}{2}\tilde{R}^{\alpha\beta\nu\rho}p_{\rho}p_{\alpha}\partial^{p}_{\beta}+p\cdot D\Sigma_{n}^{\nu\rho}D_{\rho}\biggr{)}f_{(0)}\biggr{]}\,.\end{split} (54)

From the O()O(\hbar) part, we readily find

δuf(1)=0.\delta_{u}f_{(1)}=0\,. (55)

By contracting Eq. (54) with uμu^{\mu}, we find

δuf(2)=uμpuΣμνu[Dνf(1)1p2(12R~αβνρpρpαβp+pDΣnνρDρ)f(0)].\delta_{u}f_{(2)}=-\frac{u^{\mu}}{p\cdot u}\Sigma^{u^{\prime}}_{\mu\nu}\biggl{[}D^{\nu}f_{(1)}-\frac{1}{p^{2}}\biggl{(}\frac{1}{2}\tilde{R}^{\alpha\beta\nu\rho}p_{\rho}p_{\alpha}\partial^{p}_{\beta}+p\cdot D\Sigma_{n}^{\nu\rho}D_{\rho}\biggr{)}f_{(0)}\biggr{]}\,. (56)

We can check that the above δuf(2)\delta_{u}f_{{(2)}} fulfills δ(p2)p2δuf(2)=0\delta(p^{2})p^{2}\delta_{u}f_{{(2)}}=0 and Eq. (54).

In the Wigner function (46), the frame vectors nμn^{\mu} and uμu^{\mu} are in general chosen independently. As long as f(1)f_{(1)} and f(2)f_{(2)} obey the transformation laws (51), (52), (55) and (56), however, we can always set uμ=nμu^{\mu}=n^{\mu} by redefining f(2)f_{(2)}. Then, Eq. (46) is simplified as

μ=2π[δ(p2)(pμ+ΣμνnDν)+2p2{pμQp+2pν(T[μpν]+Sαμνpα)}δ(p2)+2δ(p2)2p2{εμνρσpνDρΣnσλDλΣμνn(R~αβνρpρpαβp+2pDΣnνρDρ)}]f,\begin{split}\mathcal{R}_{\mu}&=2\pi\Biggl{[}\delta(p^{2})\bigl{(}p_{\mu}+\hbar\Sigma_{\mu\nu}^{n}D^{\nu}\bigr{)}+\frac{\hbar^{2}}{p^{2}}\Bigl{\{}-p_{\mu}Q\cdot p+2p^{\nu}\bigl{(}T_{[\mu}p_{\nu]}+S_{\alpha\mu\nu}p^{\alpha}\bigr{)}\Bigr{\}}\delta(p^{2})\\ &\qquad+\frac{\hbar^{2}\delta(p^{2})}{2p^{2}}\biggl{\{}\varepsilon_{\mu\nu\rho\sigma}p^{\nu}D^{\rho}\Sigma_{n}^{\sigma\lambda}D_{\lambda}-\Sigma_{\mu\nu}^{n}\bigl{(}{\tilde{R}}^{\alpha\beta\nu\rho}p_{\rho}p_{\alpha}\partial_{\beta}^{p}+2p\cdot D\Sigma_{n}^{\nu\rho}D_{\rho}\bigr{)}\biggr{\}}\Biggr{]}f\,,\end{split} (57)

where we define

f=f(0)+f(1)+2f(2).f=f_{(0)}+\hbar f_{(1)}+\hbar^{2}f_{(2)}\,. (58)

The transformation laws under the change of the frames nμn^{\mu} and uμu^{\mu} are helpful to identify the equilibrium distribution function. Let us first start from the classical distribution f(0)f_{(0)}, which is defined as a function of the collisional conserved quantities:

f(0)=f(0)(g(0)=βμ+βp),\displaystyle f_{(0)}=f_{(0)}(g_{(0)}=-\beta\mu+\beta\cdot p)\,, (59)
μ(βμ)=0,μβν+νβμ=0.\displaystyle\nabla_{\mu}(\beta\mu)=0\,,\quad\nabla_{\mu}\beta_{\nu}+\nabla_{\nu}\beta_{\mu}=0\,. (60)

For this f(0)f_{(0)}, the transformation law (51) yields

δnf(1)=f(0)nμpnΣμνnpρνβρ=f(0)nμpn(12Σnνρpμ14ενρμσpσ)νβρ=f(0)12(ΣnνρΣnνρ)νβρ,\begin{split}\delta_{n}f_{(1)}&=-f^{\prime}_{(0)}\frac{n_{\mu}}{p\cdot n}\Sigma_{\mu\nu}^{n^{\prime}}p^{\rho}\nabla_{\nu}\beta_{\rho}\\ &=-f^{\prime}_{(0)}\frac{n_{\mu}}{p\cdot n}\biggl{(}-\frac{1}{2}\Sigma^{\nu\rho}_{n^{\prime}}p^{\mu}-\frac{1}{4}\varepsilon^{\nu\rho\mu\sigma}p_{\sigma}\biggr{)}\nabla_{\nu}\beta_{\rho}\\ &=f^{\prime}_{(0)}\frac{1}{2}\Bigl{(}\Sigma^{\nu\rho}_{n^{\prime}}-\Sigma^{\nu\rho}_{n}\Bigr{)}\nabla_{\nu}\beta_{\rho}\,,\end{split} (61)

where we use Eq. (45) and define f(0)=df(0)(g(0))/dg(0)f^{\prime}_{(0)}=df_{(0)}(g_{(0)})/dg_{(0)}. Although the above relation identifies the frame-dependent part involved in f(1)f_{(1)} at equilibrium, the frame-independent part is still undetermined. If we set such an ambiguous part in f(1)f_{(1)} to be zero, however, we identify

f(1)=f(0)12Σnμνμβν.f_{{(1)}}=f_{(0)}^{\prime}\frac{1}{2}\Sigma_{n}^{\mu\nu}\nabla_{\mu}\beta_{\nu}\,. (62)

This is a plausible form in the sense that the spin-vorticity coupling term is correctly reproduced: f(0)+f(1)f(0)(g(0)+2Σnμνμβν)+O(2)f_{(0)}+\hbar f_{(1)}\simeq f_{(0)}(g_{(0)}+\frac{\hbar}{2}\Sigma_{n}^{\mu\nu}\nabla_{\mu}\beta_{\nu})+O(\hbar^{2}). In this case, the first order Wigner function (32) is written as

(1)eqμ=2πδ(p2)f(0)(14)εμνρσpνρβσ.\mathcal{R}^{\mu}_{{(1)}\text{eq}}=2\pi\delta(p^{2})f^{\prime}_{(0)}\biggl{(}-\frac{1}{4}\biggr{)}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}\nabla_{\rho}\beta_{\sigma}\,. (63)

This (1)eqμ\mathcal{R}^{\mu}_{{(1)}\text{eq}} fulfills the kinetic equation at O()O(\hbar) Liu:2018xip . Subsequently, with the above f(0)f_{(0)} and f(1)f_{(1)}, the transformation laws (52) and (56) lead to

δnf(2)=1p2ΣμνuDμf(0)(ΣnνρΣnνρ)pσρβσ=ΣμνuDμ[14f(0)ερσνλ(nλpnnλpn)ρβσ],δuf(2)=uμpuΣμνu[Dνf(0)12Σnρσρβσ1p2(12R~αβνρpρpαββf(0)+pDf(0)Σnνρpσρβσ)]=(ΣμνuΣμνu)Dμ(f(0)ερσνλnλ4pnρβσ),\begin{split}\delta_{n}f_{(2)}&=\frac{1}{p^{2}}\Sigma_{\mu\nu}^{u}D^{\mu}f^{\prime}_{(0)}\Bigl{(}\Sigma^{\nu\rho}_{n^{\prime}}-\Sigma^{\nu\rho}_{n}\Bigr{)}p^{\sigma}\nabla_{\rho}\beta_{\sigma}\\ &=\Sigma_{\mu\nu}^{u}D^{\mu}\biggl{[}\frac{1}{4}f^{\prime}_{(0)}\varepsilon^{\rho\sigma\nu\lambda}\biggl{(}\frac{n^{\prime}_{\lambda}}{p\cdot n^{\prime}}-\frac{n_{\lambda}}{p\cdot n}\biggr{)}\nabla_{\rho}\beta_{\sigma}\biggr{]}\,,\\ \delta_{u}f_{(2)}&=-\frac{u^{\mu}}{p\cdot u}\Sigma^{u^{\prime}}_{\mu\nu}\biggl{[}D^{\nu}f_{(0)}^{\prime}\frac{1}{2}\Sigma_{n}^{\rho\sigma}\nabla_{\rho}\beta_{\sigma}-\frac{1}{p^{2}}\biggl{(}\frac{1}{2}\tilde{R}^{\alpha\beta\nu\rho}p_{\rho}p_{\alpha}\beta_{\beta}f_{(0)}^{\prime}+p\cdot Df_{(0)}^{\prime}\Sigma^{\nu\rho}_{n}p^{\sigma}\nabla_{\rho}\beta_{\sigma}\biggr{)}\biggr{]}\\ &=\Bigl{(}\Sigma^{u^{\prime}}_{\mu\nu}-\Sigma^{u}_{\mu\nu}\Bigr{)}D^{\mu}\biggl{(}f_{(0)}^{\prime}\frac{\varepsilon^{\rho\sigma\nu\lambda}n_{\lambda}}{4p\cdot n}\nabla_{\rho}\beta_{\sigma}\biggr{)}\,,\end{split} (64)

where we employ Eq. (45). Therefore we find

f(2)=ΣμνuDμ(f(0)ενρσλ4pnnρσβλ)+ϕ(2).f_{(2)}=\Sigma^{u}_{\mu\nu}D^{\mu}\biggl{(}f^{\prime}_{(0)}\frac{\varepsilon^{\nu\rho\sigma\lambda}}{4\,p\cdot n}n_{\rho}\nabla_{\sigma}\beta_{\lambda}\biggr{)}+\phi_{(2)}\,. (65)

Here, unlike the first-order correction f(1)f_{(1)}, we explicitly keep the frame-independent ambiguity ϕ(2)\phi_{(2)}. Importantly, this ambiguous part should be taken into account for the realization of equilibrium, as we elaborate later. As shown in Appendix A, inserting the above distribution functions f(0),(1),(2)f_{{(0)},{(1)},{(2)}} into Eq. (42), we reduce the second-order correction part to

(2)eqμ=2πδ(p2)[ϕ(2)pμ+f(0)(12p2Rμαpα112p2Rpμ+23(p2)2Rαβpμpαpβ)+f(0)(124Rμαβα+112p2Rαβγμpαββpγ)+f(0)′′(124Rαβγμpαβββγ112p2Rαβγδpμpαpγβββδ14[ρβμ]pν[ρβν]+pμ4p2pν[ρβν]pσ[ρβσ])].\begin{split}\mathcal{R}^{\mu}_{{(2)}\text{eq}}&=2\pi\delta(p^{2})\Biggl{[}\phi_{(2)}p^{\mu}+f_{(0)}\biggl{(}-\frac{1}{2p^{2}}R^{\mu\alpha}p_{\alpha}-\frac{1}{12p^{2}}R\,p^{\mu}+\frac{2}{3(p^{2})^{2}}R^{\alpha\beta}p^{\mu}p_{\alpha}p_{\beta}\biggr{)}\\ &\qquad\qquad+f^{\prime}_{(0)}\biggl{(}-\frac{1}{24}R^{\mu\alpha}\beta_{\alpha}+\frac{1}{12p^{2}}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}p_{\gamma}\biggr{)}\\ &\qquad\qquad+f^{\prime\prime}_{(0)}\biggl{(}-\frac{1}{24}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}\beta_{\gamma}-\frac{1}{12p^{2}}R_{\alpha\beta\gamma\delta}p^{\mu}p^{\alpha}p^{\gamma}\beta^{\beta}\beta^{\delta}\\ &\qquad\qquad\qquad\quad-\frac{1}{4}\nabla^{[\rho}\beta^{\mu]}p^{\nu}\nabla_{[\rho}\beta_{\nu]}+\frac{p^{\mu}}{4p^{2}}p_{\nu}\nabla^{[\rho}\beta^{\nu]}p^{\sigma}\nabla_{[\rho}\beta_{\sigma]}\biggr{)}\Biggr{]}\,.\end{split} (66)

The frame-dependence here vanishes totally, as it should. Plugging this into the chiral kinetic equation (8), and after a straightforward calculation in Appendix B, we finally arrive at

0=[D(2)+P(0)]/2π=δ(p2)pDϕ(2)+f(0)δ(p2)(18βR+14p2βRαβpαpβ)+f(0)′′δ(p2)(124pRαββαββ+18Rαβμνpαββμβν)+f(0)′′′δ(p2)(124βRρσμνpμβνpρβσ)+δ(p2)(12p2Rμαpα)Dμf(0)+δ(p2)(112Rμαβα)Dμf(0)+δ(p2)(16Rαβγμpαβββγ14ρβμpνρβν)Dμf(0)′′,\begin{split}0&=\Bigl{[}D\cdot\mathcal{R}_{(2)}+P\cdot\mathcal{R}_{(0)}\Bigr{]}/2\pi\\ &=\delta(p^{2})\,p\cdot D\,\phi_{(2)}+f_{(0)}^{\prime}\delta(p^{2})\biggl{(}-\frac{1}{8}\beta\cdot\nabla R+\frac{1}{4p^{2}}\beta\cdot\nabla R^{\alpha\beta}p_{\alpha}p_{\beta}\biggr{)}\\ &\quad+f_{(0)}^{\prime\prime}\delta(p^{2})\biggl{(}-\frac{1}{24}p\cdot\nabla R^{\alpha\beta}\beta_{\alpha}\beta_{\beta}+\frac{1}{8}R^{\alpha\beta\mu\nu}p_{\alpha}\beta_{\beta}\nabla_{\mu}\beta_{\nu}\biggr{)}\\ &\quad+f_{(0)}^{\prime\prime\prime}\delta(p^{2})\biggl{(}-\frac{1}{24}\beta\cdot\nabla R_{\rho\sigma\mu\nu}p^{\mu}\beta^{\nu}p^{\rho}\beta^{\sigma}\biggr{)}+\delta(p^{2})\biggl{(}-\frac{1}{2p^{2}}R^{\mu\alpha}p_{\alpha}\biggr{)}D_{\mu}f_{(0)}\\ &\quad+\delta(p^{2})\biggl{(}\frac{1}{12}R^{\mu\alpha}\beta_{\alpha}\biggr{)}D_{\mu}f_{(0)}^{\prime}+\delta(p^{2})\biggl{(}-\frac{1}{6}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}\beta_{\gamma}-\frac{1}{4}\nabla^{\rho}\beta^{\mu}p^{\nu}\nabla_{\rho}\beta_{\nu}\biggr{)}D_{\mu}f_{(0)}^{\prime\prime}\,,\end{split} (67)

which is the equation to determine ϕ(2)\phi_{(2)}. However, there is in general no solution, as is obvious from the constraint (43); ϕ(2)\phi_{(2)} cannot have δ(p2)/p2\delta(p^{2})/p^{2} terms. In other words, the collisionless CKT has no global equilibrium solution in general background curved geometry. We note that an equilibrium distribution function with ϕ(2)=0\phi_{(2)}=0 is realized in the flat spacetime limit gμν=ημνg_{\mu\nu}=\eta_{\mu\nu}; the Killing equation μβν+νβμ=0\partial_{\mu}\beta_{\nu}+\partial_{\nu}\beta_{\mu}=0 leads to

δ(p2)(14ρβμpνρβν)μf(0)′′=δ(p2)(14ρβμ)ρμf(0)=0.\begin{split}\delta(p^{2})\biggl{(}-\frac{1}{4}\partial^{\rho}\beta^{\mu}p^{\nu}\partial_{\rho}\beta_{\nu}\biggr{)}\partial_{\mu}f_{(0)}^{\prime\prime}&=\delta(p^{2})\biggl{(}-\frac{1}{4}\partial^{\rho}\beta^{\mu}\biggr{)}\partial_{\rho}\partial_{\mu}f_{(0)}^{\prime}=0\,.\end{split} (68)

4 Stationary weak gravity

Although the general curved spacetime does not realize an equilibrium, there may exist a special geometry having a solution for Eq. (67). One of the simplest cases is the stationary and weak background gravitational field, where the metric tensor is given by

gμν=ημν+hμν,0hμν=0,|hμν|1.g_{\mu\nu}=\eta_{\mu\nu}+h_{\mu\nu}\,,\quad\partial_{0}h_{\mu\nu}=0\,,\quad|h_{\mu\nu}|\ll 1\,. (69)

In this case, the time-like Killing vector βμξμ:=δ0μ\beta^{\mu}\parallel\xi^{\mu}:=\delta^{\mu}_{0} is admitted. Then, the kinetic equation (67) is drastically reduced as

δ(p2)pD(ϕ(2)124Rαββαββf(0)′′)=0.\begin{split}\delta(p^{2})p\cdot D\biggl{(}\phi_{(2)}-\frac{1}{24}R^{\alpha\beta}\beta_{\alpha}\beta_{\beta}f_{(0)}^{\prime\prime}\biggr{)}=0\,.\end{split} (70)

Therefore, for the metric tensor (69), we identify

ϕ(2)=f(0)′′24Rαββαββ.\phi_{(2)}=\frac{f_{(0)}^{\prime\prime}}{24}R^{\alpha\beta}\beta_{\alpha}\beta_{\beta}\,. (71)

Hereafter, we call f=f(0)+f(1)+2f(2)f=f_{(0)}+\hbar f_{(1)}+\hbar^{2}f_{(2)} with Eqs. (59), (62), (65) and (71) an equilibrium distribution function. In this section, we focus on the geometry described by Eq. (69).

Let us evaluate the charge current and the symmetric energy-momentum tensor for the equilibrium distribution function. We employ the classical equilibrium state described by

f(0)=θ(βp)eg(0)+1+θ(βp)eg(0)+1,g(0)=βμ+βp,\displaystyle\displaystyle f_{(0)}=\frac{\theta(\beta\cdot p)}{e^{g_{(0)}}+1}+\frac{\theta(-\beta\cdot p)}{e^{-g_{(0)}}+1}\,,\quad g_{(0)}=-\beta\mu+\beta\cdot p\,, (72)

where βμ\beta^{\mu} is a time-like Killing vector βμ=β¯ξμ\beta^{\mu}=\bar{\beta}\xi^{\mu} with β¯=ββ/g00\bar{\beta}=\sqrt{\beta\cdot\beta/g_{00}}. Also β¯\bar{\beta} and μ¯\bar{\mu} are the global inverse temperature and chemical potential. In the flat spacetime limit, the classical charge density becomes

J(0)eqμ=2p(0)eqμ=𝒑[nF(|𝒑|μ)nF(|𝒑|+μ)],nF(z):=1eβz+1J^{\mu}_{{(0)}\text{eq}}=2\int_{p}\mathcal{R}^{\mu}_{{(0)}\text{eq}}=\int_{\boldsymbol{p}}\Bigl{[}n_{F}(|{\boldsymbol{p}}|-\mu)-n_{F}(|{\boldsymbol{p}}|+\mu)\Bigr{]}\,,\quad n_{F}(z):=\frac{1}{e^{\beta z}+1} (73)

with 𝒑=d3p/(2π)3\int_{\boldsymbol{p}}=\int d^{3}p/(2\pi)^{3}.

From Eq. (47), the equilibrium Wigner function (1)eqμ\mathcal{R}^{\mu}_{{(1)}\text{eq}} yields the CVE Chen:2015gta :

J(1)eqμ=C1ωμ,T(1)eqμν=2C2ξ(μων),\begin{split}J^{\mu}_{{(1)}\text{eq}}=C_{1}\omega^{\mu}\,,\quad T_{{(1)}\text{eq}}^{\mu\nu}=2C_{2}\xi^{(\mu}\omega^{\nu)}\,,\end{split} (74)

where the vorticity vector is introduced as ωμ=εμνρσξνρξσ/2\omega^{\mu}=\varepsilon^{\mu\nu\rho\sigma}\xi_{\nu}\nabla_{\rho}\xi_{\sigma}/2. Here the coefficients are defined as

Cn:=12π20𝑑ρρn[nF(ρμ)(1)nnF(ρ+μ)],C_{n}:=\frac{1}{2\pi^{2}}\int_{0}^{\infty}d\rho\rho^{n}\Bigl{[}n_{F}(\rho-\mu)-(-1)^{n}n_{F}(\rho+\mu)\Bigr{]}\,, (75)

and thus we find C1=μ2/(4π2)+T2/12C_{1}=\mu^{2}/(4\pi^{2})+T^{2}/12 and C2=μ3/6(π2)+μT2/6C_{2}=\mu^{3}/6(\pi^{2})+\mu T^{2}/6 (see Appendix C). The above charge current is conserved, namely, μJ(1)eqμ=0\nabla_{\mu}J_{{(1)}\text{eq}}^{\mu}=0. This reflects the absence of the gravitational contribution to the U(1)U(1) anomaly in the CKT up to O()O(\hbar). Besides, the energy-momentum conservation holds, as we check μT(1)eqμν=0\nabla_{\mu}T_{{(1)}\text{eq}}^{\mu\nu}=0.

From Eq. (66) and (71), the second order equilibrium Wigner function reads

(2)eqμ=2πδ(p2)[f(0)(12p2Rμαpα112p2Rpμ+23(p2)2Rαβpμpαpβ)+f(0)(124Rμαβα+112p2Rαβγμpαββpγ)+f(0)′′(124Rαβγμpαβββγ112p2Rαβγδpμpαpγβββδ+124Rαββαββ)],\begin{split}\mathcal{R}^{\mu}_{{(2)}\text{eq}}&=2\pi\delta(p^{2})\Biggl{[}f_{(0)}\biggl{(}-\frac{1}{2p^{2}}R^{\mu\alpha}p_{\alpha}-\frac{1}{12p^{2}}R\,p^{\mu}+\frac{2}{3(p^{2})^{2}}R^{\alpha\beta}p^{\mu}p_{\alpha}p_{\beta}\biggr{)}\\ &\qquad\qquad+f^{\prime}_{(0)}\biggl{(}-\frac{1}{24}R^{\mu\alpha}\beta_{\alpha}+\frac{1}{12p^{2}}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}p_{\gamma}\biggr{)}\\ &\qquad\qquad+f^{\prime\prime}_{(0)}\biggl{(}-\frac{1}{24}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}\beta_{\gamma}-\frac{1}{12p^{2}}R_{\alpha\beta\gamma\delta}p^{\mu}p^{\alpha}p^{\gamma}\beta^{\beta}\beta^{\delta}+\frac{1}{24}R^{\alpha\beta}\beta_{\alpha}\beta_{\beta}\biggr{)}\Biggr{]}\,,\end{split} (76)

where R=gμνRμνR=g^{\mu\nu}R_{\mu\nu} is the Ricci scalar. Here we dropped the terms including ρβν\nabla_{\rho}\beta_{\nu} because they are of order O(hμν2)O(h_{\mu\nu}^{2}). In the momentum integral, the 1/p21/p^{2} terms can be rewritten as

pδ(p2)p2pμF(p)=p12δ(p2)pμF(p),\displaystyle\displaystyle\int_{p}\frac{\delta(p^{2})}{p^{2}}p^{\mu}F(p)=\int_{p}\frac{1}{2}\delta(p^{2})\partial^{\mu}_{p}F(p)\,, (77)

which follows from δ(x)=δ(x)/x\delta^{\prime}(x)=-\delta(x)/x. The integral with 1/(p2)21/(p^{2})^{2} is also computed in a similar manner with δ′′(x)=2δ(x)/x2\delta^{\prime\prime}(x)=2\delta(x)/x^{2}. With the help of several formulas in Appendix C, we eventually derive

J(2)eqμ=C0[112Rμαξα124ξμR+16ξμRαβξαξβ],T(2)eqμν=C1[112Rμν112Rξμξν+124Rημν16Rα(μξν)ξα+16Rαβξαξβ(4ξμξνημν)+16Rμανβξαξβ]\begin{split}J^{\mu}_{{(2)}\text{eq}}&=C_{0}\biggl{[}\frac{1}{12}{R^{\mu}}_{\alpha}\xi^{\alpha}-\frac{1}{24}\xi^{\mu}R+\frac{1}{6}\xi^{\mu}R_{\alpha\beta}\xi^{\alpha}\xi^{\beta}\biggr{]}\,,\\ T_{{(2)}\text{eq}}^{\mu\nu}&=C_{1}\biggl{[}-\frac{1}{12}R^{\mu\nu}-\frac{1}{12}R\xi^{\mu}\xi^{\nu}+\frac{1}{24}R\eta^{\mu\nu}\\ &\qquad\quad-\frac{1}{6}R^{\alpha(\mu}\xi^{\nu)}\xi_{\alpha}+\frac{1}{6}R^{\alpha\beta}\xi_{\alpha}\xi_{\beta}(4\xi^{\mu}\xi^{\nu}-\eta^{\mu\nu})+\frac{1}{6}R^{\mu\alpha\nu\beta}\xi_{\alpha}\xi_{\beta}\biggr{]}\end{split} (78)

with C0=μ/(2π2)C_{0}=\mu/(2\pi^{2}). Therefore, based on the CKT, we reveal that the nonvanishing fermion transport is induced by the background gravitational field even in equilibrium. This implies that these phenomena are nondissipative, as so are the CME, and CVE. This is one of the main findings of this paper. We can also derive the same current J(2)eqμJ^{\mu}_{{(2)}\text{eq}} from the diagrammatic computation (see Appendix D.1) and with the Riemann normal coordinate expansion (see Appendix D.2). It is worthwhile to mention that g0ig_{0i} enters in Eqs. (74) and (78) only through the field strength fij=ig0jjg0if_{ij}=\partial_{i}g_{0j}-\partial_{j}g_{0i}. This is a consequence of the Kaluza-Klein gauge symmetry Banerjee:2012iz .

For the left-handed Weyl fermion, J(2)eqμJ^{\mu}_{{(2)}\text{eq}} and T(2)eqμνT^{\mu\nu}_{{(2)}\text{eq}} are written as the same form, while the sign of J(1)eqμJ^{\mu}_{{(1)}\text{eq}} and T(1)eqμνT^{\mu\nu}_{{(1)}\text{eq}} flipped; the former does not involve εμνρσ\varepsilon^{\mu\nu\rho\sigma} while the latter does. As a result, the vector and axial parts are written as

J(2)eqV/Aμ=C0,V/A[112Rμαξα124ξμR+16ξμRαβξαξβ],T(2)eqV/Aμν=C1,V/A[112Rμν112Rξμξν+124Rημν16Rα(μξν)ξα+16Rαβξαξβ(4ξμξνημν)+16Rμανβξαξβ]\begin{split}J^{\mu}_{{(2)}\text{eq}\,V/A}&=C_{0,V/A}\biggl{[}\frac{1}{12}{R^{\mu}}_{\alpha}\xi^{\alpha}-\frac{1}{24}\xi^{\mu}R+\frac{1}{6}\xi^{\mu}R_{\alpha\beta}\xi^{\alpha}\xi^{\beta}\biggr{]}\,,\\ T_{{(2)}\text{eq}\,V/A}^{\mu\nu}&=C_{1,V/A}\biggl{[}-\frac{1}{12}R^{\mu\nu}-\frac{1}{12}R\xi^{\mu}\xi^{\nu}+\frac{1}{24}R\eta^{\mu\nu}\\ &\qquad\quad-\frac{1}{6}R^{\alpha(\mu}\xi^{\nu)}\xi_{\alpha}+\frac{1}{6}R^{\alpha\beta}\xi_{\alpha}\xi_{\beta}(4\xi^{\mu}\xi^{\nu}-\eta^{\mu\nu})+\frac{1}{6}R^{\mu\alpha\nu\beta}\xi_{\alpha}\xi_{\beta}\biggr{]}\end{split} (79)

where C0,V/A=μV/A/π2C_{0,\,V/A}=\mu_{V/A}/\pi^{2}, C1,V=(μV2+μA2)/2π2+T2/6C_{1,\,V}=(\mu_{V}^{2}+\mu_{A}^{2})/2\pi^{2}+T^{2}/6 and C1,A=μVμA/π2C_{1,\,A}=\mu_{V}\mu_{A}/\pi^{2}, with μV\mu_{V} and μA\mu_{A} being the vector and chiral chemical potential, respectively. In Sec. 7, we argue the novelty and some implications of Eq. (78) and (79).

5 Dynamical weak gravity

While so far we have focused on the equilibrium state, this section is dedicated to discuss the dynamical response from the time-dependent gravity. Specifically, we consider a plane-wave weak background gravitational field:

gμν=ημν+hμν,hμνeikx,|hμν|1,g_{\mu\nu}=\eta_{\mu\nu}+h_{\mu\nu}\,,\quad h_{\mu\nu}\sim e^{-ik\cdot x}\,,\quad|h_{\mu\nu}|\ll 1\,, (80)

where kμ=(k0,𝒌)k^{\mu}=(k_{0},{\boldsymbol{k}}) is the momentum of the gravitational field. Let us look for the perturbative distribution function represented as the following form:

f=fflat+f~,fflat=θ(p0)eβ(p0μ)+1+θ(p0)eβ(p0μ)+1,f~=f~(0)+f~(1)+2f~(2).f=f_{\text{\text{flat}}}+\tilde{f}\,,\quad f_{\text{flat}}=\frac{\theta(p_{0})}{e^{\beta(p_{0}-\mu)}+1}+\frac{\theta(-p_{0})}{e^{-\beta(p_{0}-\mu)}+1}\,,\quad\tilde{f}=\tilde{f}_{(0)}+\hbar\tilde{f}_{(1)}+\hbar^{2}\tilde{f}_{(2)}\,. (81)

Here β\beta and μ\mu are constant, and thus fflatf_{\text{flat}} is the static and homogeneous solution of the collisionless Boltzmann equation (27) for hμν=0h_{\mu\nu}=0. We define f~\tilde{f} as the fluctuation around fflatf_{\text{flat}}. For hμνeikxh_{\mu\nu}\sim e^{-ik\cdot x}, we may employ the anzatz f~eikx\tilde{f}\sim e^{-ik\cdot x}. For simplicity, we further assume μnν=0\partial_{\mu}n^{\nu}=0.

We first compute the classical and leading order parts. Plugging the general form of (0),(1)μ\mathcal{R}^{\mu}_{{(0)},{(1)}} in Eqs. (26) and (32) into Eq. (8), we write down the kinetic equation as

δ(p2)[pD+(DμΣnμν)Dν2ΣnμνRαβμνpαpβ]f=0.\begin{split}\delta(p^{2})\biggl{[}p\cdot D+\hbar(D_{\mu}\Sigma_{n}^{\mu\nu})D_{\nu}-\frac{\hbar}{2}\Sigma^{\mu\nu}_{n}R_{\alpha\beta\mu\nu}p^{\alpha}\partial_{p}^{\beta}\biggr{]}f=0\,.\end{split} (82)

Expanding the above equation in terms of hμνh_{\mu\nu} and utilizing μfflat=0\partial_{\mu}f_{\text{flat}}=0, we obtain

pf~+[Γμνρpμpρβν2ΣnμνRαβμνpαββ]fflat=0,p\cdot\partial\tilde{f}+\biggl{[}\Gamma^{\rho}_{\mu\nu}p^{\mu}p_{\rho}\beta^{\nu}-\frac{\hbar}{2}\Sigma^{\mu\nu}_{n}R_{\alpha\beta\mu\nu}p^{\alpha}\beta^{\beta}\biggr{]}f^{\prime}_{\text{flat}}=0\,, (83)

where we denote βμ=βξμ=βδ0μ\beta^{\mu}=\beta\xi^{\mu}=\beta\delta^{\mu}_{0}. Note that after the weak gravitational field expansion, all indices are raised and lowered by ημν\eta_{\mu\nu} and the inner products are defined as AB=ημνAμBνA\cdot B=\eta_{\mu\nu}A^{\mu}B^{\nu} and A2=ημνAμAνA^{2}=\eta_{\mu\nu}A^{\mu}A^{\nu}. Especially, to get the above equation, we have taken the following replacement:

pμ\displaystyle\displaystyle p^{\mu} \displaystyle\to gμνpνpμhμνpν,\displaystyle g^{\mu\nu}p_{\nu}\simeq p^{\mu}-h^{\mu\nu}p_{\nu}\,, (84)
δ(p2)\displaystyle\displaystyle\delta(p^{2}) \displaystyle\to δ(gμνpμpν)δ(p2)(1+1p2hμνpμpν),\displaystyle\delta(g^{\mu\nu}p_{\mu}p_{\nu})\simeq\delta(p^{2})\biggl{(}1+\frac{1}{p^{2}}h^{\mu\nu}p_{\mu}p_{\nu}\biggr{)}\,, (85)

which follow from gμνημνhμνg^{\mu\nu}\simeq\eta^{\mu\nu}-h^{\mu\nu} and δ(x)=δ(x)/x\delta^{\prime}(x)=-\delta(x)/x. For hμνeikxh_{\mu\nu}\sim e^{-ik\cdot x}, the fluctuations f~(0),(1)eikx\tilde{f}_{{(0)},{(1)}}\sim e^{-ik\cdot x} are found to be

f~(0)=1ikpΓμνρpμpρβνfflat,\displaystyle\displaystyle\tilde{f}_{{(0)}}=\frac{1}{ik\cdot p}\Gamma^{\rho}_{\mu\nu}p^{\mu}p_{\rho}\beta^{\nu}f^{\prime}_{\text{flat}}\,, (86)
f~(1)=12ikpΣnμνRαβμνpαββfflat,\displaystyle\displaystyle\tilde{f}_{{(1)}}=-\frac{1}{2ik\cdot p}\Sigma^{\mu\nu}_{n}{R}_{\alpha\beta\mu\nu}p^{\alpha}\beta^{\beta}f^{\prime}_{\text{flat}}\,, (87)

with the linearized Christoffel symbol and Riemann tensor being

Γμλρi2(kμhλρ+kλhμρkρhμλ),Rλμνρ(i)22(kνkλhμρkνkρhμλkμkλhνρ+kμkρhνλ).\begin{split}{\Gamma}_{\mu\lambda}^{\rho}&\simeq\frac{-i}{2}(k_{\mu}h_{\lambda}^{\rho}+k_{\lambda}h_{\mu}^{\rho}-k^{\rho}h_{\mu\lambda})\,,\quad\\ {R}^{\rho}_{~{}\lambda\mu\nu}&\simeq\frac{(-i)^{2}}{2}(k_{\nu}k_{\lambda}h_{\mu}^{\rho}-k_{\nu}k^{\rho}h_{\mu\lambda}-k_{\mu}k_{\lambda}h_{\nu}^{\rho}+k_{\mu}k^{\rho}h_{\nu\lambda})\,.\end{split} (88)

Plugging the above distribution functions into Eqs. (26) and (32), we get the Wigner function. It is here informative to decompose the Wigner function into the terms that involve the kk-dependent pole in the denominator, and the others. As we show in Sec. 6, the momentum integrals of the former vanishes in the static limit k0/|𝒌|0k_{0}/|{\boldsymbol{k}}|\to 0, while those of the latter survives. In this sense, we denote such a (non)static part as (non)stμ\mathcal{R}^{\mu}_{\mathrm{(non)}\text{st}}. We note that the static part stμ\mathcal{R}^{\mu}_{\text{st}} reproduces the equilibrium Wigner function eqμ\mathcal{R}^{\mu}_{\text{eq}} in the previous section, as we show later.

For the classical O(0)O(\hbar^{0}) part, Eq. (86) leads to (0)μ=(0)stμ+(0)nonstμ\mathcal{R}^{\mu}_{{(0)}}=\mathcal{R}^{\mu}_{{(0)}\text{st}}+\mathcal{R}^{\mu}_{{(0)}\text{nonst}} with

(0)stμ\displaystyle\mathcal{R}^{\mu}_{{(0)}\text{st}} =\displaystyle= 2πδ(p2)[pμ(1+1p2hαβpαpβ)fflathμνpνfflat],\displaystyle 2\pi\delta(p^{2})\biggl{[}p^{\mu}\biggl{(}1+\frac{1}{p^{2}}h^{\alpha\beta}p_{\alpha}p_{\beta}\biggr{)}f_{\text{flat}}-h^{\mu\nu}p_{\nu}f_{\text{flat}}\biggr{]}\,, (89)
(0)nonstμ\displaystyle\mathcal{R}^{\mu}_{{(0)}\text{nonst}} =\displaystyle= 2πδ(p2)1ikpΓλνρpμpλpρβνfflat,\displaystyle 2\pi\delta(p^{2})\frac{1}{ik\cdot p}\Gamma^{\rho}_{\lambda\nu}p^{\mu}p^{\lambda}p_{\rho}\beta^{\nu}f_{\text{flat}}^{\prime}\,, (90)

where we use gμνημνhμνg^{\mu\nu}\simeq\eta^{\mu\nu}-h^{\mu\nu} and δ(gαβpαpβ)δ(p2)(1+hαβpαpβ/p2)\delta(g^{\alpha\beta}p_{\alpha}p_{\beta})\simeq\delta(p^{2})(1+h^{\alpha\beta}p_{\alpha}p_{\beta}/p^{2}). Similarly, Eqs. (86) and (87) yield to the O()O(\hbar) part as

(1)μ=2πδ(p2)(pμf~(1)+ΣnμνDν(fflat+f~(0)))=2πδ(p2)41ikpεμηνλpηRρσλνpρβσfflat,\begin{split}\mathcal{R}^{\mu}_{(1)}&=2\pi\delta(p^{2})\Bigl{(}p^{\mu}\tilde{f}_{{(1)}}+\Sigma_{n}^{\mu\nu}D_{\nu}(f_{\text{flat}}+\tilde{f}_{(0)})\Bigr{)}\\ &=-\frac{2\pi\delta(p^{2})}{4}\frac{1}{ik\cdot p}\varepsilon^{\mu\eta\nu\lambda}p_{\eta}R_{\rho\sigma\lambda\nu}p^{\rho}\beta^{\sigma}f_{\text{flat}}^{\prime}\,,\end{split} (91)

where we used Eq. (45) to remove Σnμν\Sigma_{n}^{\mu\nu}. We again stress that while f~(1)\tilde{f}_{(1)} is the frame-dependent, the Wigner function (1)μ\mathcal{R}^{\mu}_{(1)} is irrelevant to the frame. Then, the Wigner function is represented as (1)μ=(1)stμ+(1)nonstμ\mathcal{R}^{\mu}_{(1)}=\mathcal{R}^{\mu}_{{(1)}\text{st}}+\mathcal{R}^{\mu}_{{(1)}\text{nonst}} with

(1)stμ\displaystyle\displaystyle\mathcal{R}^{\mu}_{{(1)}\text{st}} =\displaystyle= 2πδ(p2)4εμνρσpν(ikρ)hσλβλfflat,\displaystyle-\frac{2\pi\delta(p^{2})}{4}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}(-ik_{\rho})h_{\sigma\lambda}\beta^{\lambda}f_{\text{flat}}^{\prime}\,, (92)
(1)nonstμ\displaystyle\displaystyle\mathcal{R}^{\mu}_{{(1)}\text{nonst}} =\displaystyle= 2πδ(p2)4εμνρσpν(ikρ)kβkphσλpλfflat.\displaystyle\frac{2\pi\delta(p^{2})}{4}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}(-ik_{\rho})\frac{k\cdot\beta}{k\cdot p}h_{\sigma\lambda}p^{\lambda}f_{\text{flat}}^{\prime}\,. (93)

We observe that the above (1)stμ\mathcal{R}^{\mu}_{{(1)}\text{st}} is consistent with the equilibrium Wigner function (1)eqμ\mathcal{R}^{\mu}_{{(1)}\text{eq}} in Eq. (63).

Applying the totally same manner to the O(2)O(\hbar^{2}) parts, we obtain f~2\tilde{f}_{2} and eventually the Wigner function as (2)μ=(2)stμ+(2)nonstμ\mathcal{R}^{\mu}_{{(2)}}=\mathcal{R}^{\mu}_{{(2)}\text{st}}+\mathcal{R}^{\mu}_{{(2)}\text{nonst}} with

(2)stμ=2πδ(p2)[14p2Rαβμνpνpαββfflat+14p2Rβνpμpνββfflat14Rβμββfflat+124pμRαββαββfflat′′]+2πp2[pμQp+2pν(T[μpν]+Sαμνpα)]δ(p2)fflat,\begin{split}\mathcal{R}^{\mu}_{{(2)}\text{st}}&=2\pi\delta(p^{2})\biggl{[}-\frac{1}{4p^{2}}{R_{\alpha\beta}}^{\mu\nu}p_{\nu}p^{\alpha}\beta^{\beta}f_{\text{flat}}^{\prime}+\frac{1}{4p^{2}}{R_{\beta}}^{\nu}p^{\mu}p_{\nu}\beta^{\beta}f_{\text{flat}}^{\prime}-\frac{1}{4}{R_{\beta}}^{\mu}\beta^{\beta}f_{\text{flat}}^{\prime}\\ &\quad+\frac{1}{24}p^{\mu}R_{\alpha\beta}\beta^{\alpha}\beta^{\beta}f_{\text{flat}}^{\prime\prime}\biggr{]}+\frac{2\pi}{p^{2}}\Bigl{[}-p^{\mu}Q\cdot p+2p_{\nu}\bigl{(}T^{[\mu}p^{\nu]}+S^{\alpha\mu\nu}p_{\alpha}\bigr{)}\Bigr{]}\delta(p^{2})f_{\text{flat}}\,,\end{split} (94)
(2)nonstμ=2πδ(p2)kβkp[14p2Rανpμpνpαfflat+14Rαμpαfflat+124pμpηpρβνβσRρσηνfflat′′′],\begin{split}\mathcal{R}^{\mu}_{{(2)}\text{nonst}}&=2\pi\delta(p^{2})\frac{k\cdot\beta}{k\cdot p}\biggl{[}-\frac{1}{4p^{2}}{R_{\alpha}}^{\nu}p^{\mu}p_{\nu}p^{\alpha}f_{\text{flat}}^{\prime}+\frac{1}{4}{R_{\alpha}}^{\mu}p^{\alpha}f_{\text{flat}}^{\prime}+\frac{1}{24}p^{\mu}p^{\eta}p^{\rho}\beta^{\nu}\beta^{\sigma}R_{\rho\sigma\eta\nu}f_{\text{flat}}^{\prime\prime\prime}\biggr{]}\,,\end{split} (95)

for which the precise derivation is shown in Appendix F. Again (2)stμ\mathcal{R}^{\mu}_{{(2)}\text{st}} is the same as (2)eqμ\mathcal{R}^{\mu}_{{(2)}\text{eq}} in Eq. (66) up to O(hμν)O(h_{\mu\nu}).

6 Dynamical response

In the following discussion, we evaluate the charge current JμJ^{\mu} and the energy-momentum tensor TμνT^{\mu\nu} in Eq. (47) with Eqs. (94), and (95). As the Wigner function μ\mathcal{R}^{\mu} is decomposed into the static (kk-independent) and nonstatic (kk-dependent) part, so are JμJ^{\mu} and TμνT^{\mu\nu}, that is, Jμ=Jstμ+JnonstμJ^{\mu}=J^{\mu}_{\text{st}}+J^{\mu}_{\text{nonst}} and Tμν=Tstμν+TnonstμνT^{\mu\nu}=T^{\mu\nu}_{\text{st}}+T^{\mu\nu}_{\text{nonst}}. The static part is calculated in the same manner as before. For instance, using the integral formulas in Appendix C, the momentum integrals of the classical contribution (89) yield

J(0)stμ=C2ξμ(12ξαξβhαβ),T(0)stμν=C3[43ξμξν13ημν+(13ημαηνβ4ξμξνξαξβ+23ημνξαξβ)hαβ],\begin{split}J^{\mu}_{{(0)}\text{st}}&=C_{2}\xi^{\mu}(1-2\xi^{\alpha}\xi^{\beta}h_{\alpha\beta})\,,\\ T^{\mu\nu}_{{(0)}\text{st}}&=C_{3}\biggl{[}\frac{4}{3}\xi^{\mu}\xi^{\nu}-\frac{1}{3}\eta^{\mu\nu}+\biggl{(}\frac{1}{3}\eta^{\mu\alpha}\eta^{\nu\beta}-4\xi^{\mu}\xi^{\nu}\xi^{\alpha}\xi^{\beta}+\frac{2}{3}\eta^{\mu\nu}\xi^{\alpha}\xi^{\beta}\biggr{)}h_{\alpha\beta}\biggr{]},\end{split} (96)

with C3=μ4/24π2+μ2T2/12C_{3}=\mu^{4}/24\pi^{2}+\mu^{2}T^{2}/12. Here we use (g)1/21hμμ/2(-g)^{-1/2}\simeq 1-h^{\mu}_{\mu}/2, and perform the integration by part.

For the nonstatic part, it is helpful to additionally prepare the following tensor (scalar for n=0n=0) function:

Ij1jn(x):=xdΩ4πp^j1p^jnx𝒌^𝒑^,\displaystyle I^{j_{1}\cdots j_{n}}(x):=x\int\frac{d\Omega}{4\pi}\frac{{\hat{p}}^{j_{1}}\cdots{\hat{p}}^{j_{n}}}{x-{\hat{\boldsymbol{k}}}\cdot{\hat{\boldsymbol{p}}}}\,, (97)

where we define x:=k0/|𝒌|x:=k_{0}/|{\boldsymbol{k}}| and the integral is over the solid angle of 𝒑{\boldsymbol{p}}. The evaluations of Ij1jn(x)I^{j_{1}\cdots j_{n}}(x)’s are summarized in Appendix G. Here xx in the denominators is understood to involve the positive infinitesimal imaginary part +iη+i\eta. The nonstatic part of the classical charge current is from Eq. (90) evaluated as

J(0)nonstμ=kβhλρp2πδ(p2)1kppμpλpρfflat=32C2[ξμ(h00I+hjkIjk+2h0jIj)+δiμ(h00Ii+hjkIijk+2h0jIij)].\begin{split}J^{\mu}_{{(0)}\text{nonst}}&=-k\cdot\beta h_{\lambda\rho}\int_{p}2\pi\delta(p^{2})\frac{1}{k\cdot p}p^{\mu}p^{\lambda}p^{\rho}f_{\text{flat}}^{\prime}\\ &=\frac{3}{2}C_{2}\biggl{[}\xi^{\mu}\Bigl{(}h_{00}I+h_{jk}I^{jk}+2h_{0j}I^{j}\Bigr{)}+\delta^{\mu}_{i}\Bigl{(}h_{00}I^{i}+h_{jk}I^{ijk}+2h_{0j}I^{ij}\Bigr{)}\biggr{]}\,.\end{split} (98)

To obtain the above second line, we utilized Ij1jn(x)=(1)nIj1jn(x)I^{j_{1}\cdots j_{n}}(-x)=(-1)^{n}I^{j_{1}\cdots j_{n}}(x) and

2kβdp02π2πδ(p2)(p0)nkpdmdp0mfflat=β|𝒑|n2[xx𝒌^𝒑^+iηnF(m)(|𝒑|μ)+(1)n+m+1xx𝒌^𝒑^iηnF(m)(|𝒑|+μ)]\begin{split}&2k\cdot\beta\int\frac{dp_{0}}{2\pi}2\pi\delta(p^{2})\frac{(p_{0})^{n}}{k\cdot p}\frac{d^{m}}{dp_{0}^{m}}f_{\text{flat}}\\ &=\beta|{\boldsymbol{p}}|^{n-2}\Biggl{[}\frac{x}{x-{\hat{\boldsymbol{k}}}\cdot{\hat{\boldsymbol{p}}}+i\eta}n_{F}^{(m)}(|{\boldsymbol{p}}|-\mu)+(-1)^{n+m+1}\frac{-x}{-x-{\hat{\boldsymbol{k}}}\cdot{\hat{\boldsymbol{p}}}-i\eta}n_{F}^{(m)}(|{\boldsymbol{p}}|+\mu)\Biggr{]}\end{split} (99)

with nF(y)=(eβy+1)1n_{F}({y})=(e^{\beta{y}}+1)^{-1} and nF(m)(y):=dmnF(y)/dymn_{F}^{(m)}(y):={d^{m}n_{F}(y)/dy^{m}}. Besides, the nonstatic part of the classical energy-momentum tensor is computed as

T(0)nonstμν=kβhρλp2πδ(p2)1kppμpνpρpλfflat=2C3[ξμξν(h00I+2h0jIj+hjkIjk)+2δi(μδ0ν)(h00Ii+2h0jIij+hjkIijk)+δiμδjν(h00Iij+2h0kIijk+hklIijkl)].\begin{split}T^{\mu\nu}_{{(0)}\text{nonst}}&=-k\cdot\beta h_{\rho\lambda}\int_{p}2\pi\delta(p^{2})\frac{1}{k\cdot p}p^{\mu}p^{\nu}p^{\rho}p^{\lambda}f_{\text{flat}}^{\prime}\\ &=2C_{3}\biggl{[}\xi^{\mu}\xi^{\nu}\Bigl{(}h_{00}I+2h_{0j}I^{j}+h_{jk}I^{jk}\Bigr{)}+2\delta^{(\mu}_{i}\delta^{\nu)}_{0}\Bigl{(}h_{00}I^{i}+2h_{0j}I^{ij}+h_{jk}I^{ijk}\Bigr{)}\\ &\qquad\qquad+\delta^{\mu}_{i}\delta^{\nu}_{j}\Bigl{(}h_{00}I^{ij}+2h_{0k}I^{ijk}+h_{kl}I^{ijkl}\Bigr{)}\biggr{]}\,.\end{split} (100)

It is worthwhile to notice several properties of the above Ij1jn(x)I_{j_{1}\cdots j_{n}}(x). First we find the following relations:

k0I+kkIk=k0,\displaystyle\displaystyle k_{0}I+k_{k}I^{k}=k_{0}\,, (101)
k0Ik+kjIjk=0,\displaystyle\displaystyle k_{0}I^{k}+k_{j}I^{jk}=0\,, (102)
k0Ijk+kiIijk=k0δjk3,\displaystyle\displaystyle k_{0}I^{jk}+k_{i}I^{ijk}=k_{0}\frac{\delta^{jk}}{3}\,, (103)
k0Ijkl+kiIijkl=0.\displaystyle\displaystyle k_{0}I^{jkl}+k_{i}I^{ijkl}=0\,. (104)

From these, we can show the charge current and energy-momentum conservation for arbitrary kμk^{\mu}:

μJ(0)μ=μ(J(0)stμ+J(0)nonstμ)=0,μT(0)μν=μ(T(0)stμν+T(0)nonstμν)=0.\begin{split}\nabla_{\mu}J^{\mu}_{{(0)}}&=\nabla_{\mu}(J^{\mu}_{{(0)}\text{st}}+J^{\mu}_{{(0)}\text{nonst}})=0\,,\\ \nabla_{\mu}T^{\mu\nu}_{(0)}&=\nabla_{\mu}(T^{\mu\nu}_{{(0)}\text{st}}+T^{\mu\nu}_{{(0)}\text{nonst}})=0\,.\end{split} (105)

Second, we check that Ij1jn(x)I^{j_{1}\cdots j_{n}}(x) fulfills another type of relations:

I+Ijj=0,\displaystyle I+{I^{j}}_{j}=0\,, (106)
Ik+Ijjk=0,\displaystyle I^{k}+{I_{j}}^{jk}=0\,, (107)
Ikl+Ijjkl=0.\displaystyle I^{kl}+{I_{j}}^{jkl}=0\,. (108)

These bring the dilatation current conservation for arbitrary kμk^{\mu}:

gμνT(0)μν=gμν(T(0)stμν+T(0)nonstμν)=0.g_{\mu\nu}T^{\mu\nu}_{(0)}=g_{\mu\nu}(T^{\mu\nu}_{{(0)}\text{st}}+T^{\mu\nu}_{{(0)}\text{nonst}})=0\,. (109)

As a particular case, we consider the dynamical limit x=k0/|𝒌|1x=k_{0}/|{\boldsymbol{k}}|\gg 1. We expand J(0)nonstμJ^{\mu}_{{(0)}\text{nonst}} and T(0)nonstμνT^{\mu\nu}_{{(0)}\text{nonst}} in terms of 1/x1/x, with the asymptotic forms of Ij1jn(x)I_{j_{1}\cdots j_{n}}(x)’s, which are derived in Eqs. (215)-(219). For later convenience, we here define the total charge current and energy-momentum tensor in the dynamical limit, as follows:

Jdynμ:=Jstμ+Jnonstμ|x,Tdynμν:=Tstμν+Tnonstμν|x.J^{\mu}_{\text{dyn}}:=J^{\mu}_{\text{st}}+J^{\mu}_{\text{nonst}}\bigl{|}_{x\to\infty}\,,\quad T^{\mu\nu}_{\text{dyn}}:=T^{\mu\nu}_{\text{st}}+T^{\mu\nu}_{\text{nonst}}\bigl{|}_{x\to\infty}\,. (110)

Their classical contributions hence become

J(0)dynμ=C2[ξμ(112hλλ)δiμh0i],T(0)dynμν=C3[13(4ξμξνημν)+(35ημαηνβ+215ημνηαβ+125ξμξνξαξβ45ηαβξμξν215ξαξβημν165ηα(μξν)ξβ)hαβ].\begin{split}J^{\mu}_{{(0)}\text{dyn}}&=C_{2}\biggl{[}\xi^{\mu}\biggl{(}1-\frac{1}{2}{h^{\lambda}}_{\lambda}\biggr{)}-\delta^{\mu}_{i}h^{i}_{0}\biggr{]}\,,\\ T^{\mu\nu}_{{(0)}\text{dyn}}&=C_{3}\biggl{[}\frac{1}{3}\Bigl{(}4\xi^{\mu}\xi^{\nu}-\eta^{\mu\nu}\Bigr{)}+\biggl{(}\frac{3}{5}\eta^{\mu\alpha}\eta^{\nu\beta}+\frac{2}{15}\eta^{\mu\nu}\eta^{\alpha\beta}\\ &\qquad\quad+\frac{12}{5}\xi^{\mu}\xi^{\nu}\xi^{\alpha}\xi^{\beta}-\frac{4}{5}\eta^{\alpha\beta}\xi^{\mu}\xi^{\nu}-\frac{2}{15}\xi^{\alpha}\xi^{\beta}\eta^{\mu\nu}-\frac{16}{5}\eta^{\alpha(\mu}\xi^{\nu)}\xi^{\beta}\biggr{)}h_{\alpha\beta}\biggr{]}\,.\end{split} (111)

Let us also calculate quantum corrections. At O()O(\hbar), the Wigner function (93) leads to

J(1)nonstμ=C1ωμI12C1εμνρσ(ikρ)hσλ[(ξνδλk+δνkξλ)Ik+δνjδλkIjk],T(1)nonstμν=3C22ξ(μων)I3C24[ξ(μεν)ηρσ(ikρ)hσλ((δηkξλ+δλkξη)Ik+δηjδλkIjk)+ηi(μεν)ηρσ(ikρ)hσλ(ξηξλIi+(δηkξλ+δλkξη)Iik+δηjδλkIijk)],\begin{split}J^{\mu}_{{(1)}\text{nonst}}&=-C_{1}\omega^{\mu}I-\frac{1}{2}C_{1}\varepsilon^{\mu\nu\rho\sigma}(-ik_{\rho})h_{\sigma}^{\lambda}\biggl{[}\Bigl{(}\xi_{\nu}\delta^{k}_{\lambda}+\delta^{k}_{\nu}\xi_{\lambda}\Bigr{)}I_{k}+\delta^{j}_{\nu}\delta^{k}_{\lambda}I_{jk}\biggr{]}\,,\\ T^{\mu\nu}_{{(1)}\text{nonst}}&=-\frac{3C_{2}}{2}\xi^{(\mu}\omega^{\nu)}I-\frac{3C_{2}}{4}\biggl{[}\xi^{(\mu}\varepsilon^{\nu)\eta\rho\sigma}(-ik_{\rho})h_{\sigma}^{\lambda}\Bigl{(}(\delta_{\eta}^{k}\xi_{\lambda}+\delta_{\lambda}^{k}\xi_{\eta})I_{k}+\delta_{\eta}^{j}\delta_{\lambda}^{k}I_{jk}\Bigr{)}\\ &\qquad+\eta^{i(\mu}\varepsilon^{\nu)\eta\rho\sigma}(-ik_{\rho})h_{\sigma}^{\lambda}\Bigl{(}\xi_{\eta}\xi_{\lambda}I_{i}+(\delta_{\eta}^{k}\xi_{\lambda}+\delta_{\lambda}^{k}\xi_{\eta})I_{ik}+\delta_{\eta}^{j}\delta_{\lambda}^{k}I_{ijk}\Bigr{)}\biggr{]}\,,\end{split} (112)

where the vorticity is linearized as

ωμ=12εμνρσξνρξσ12εμνρσξν(ikρ)hσλξλ.\omega^{\mu}=\frac{1}{2}\varepsilon^{\mu\nu\rho\sigma}\xi_{\nu}\nabla_{\rho}\xi_{\sigma}\simeq\frac{1}{2}\varepsilon^{\mu\nu\rho\sigma}\xi_{\nu}(-ik_{\rho})h_{\sigma\lambda}\xi^{\lambda}\,. (113)

In particular, taking the dynamical limit, we find

J(1)nonstμ|x=C1ωμ,T(1)nonstμν|x=2C2ξ(μων)+C25δk(μεν)0lm(ikl)hmk.\begin{split}J^{\mu}_{{(1)}\text{nonst}}\bigl{|}_{x\to\infty}&=-C_{1}\omega^{\mu}\,,\\ T^{\mu\nu}_{{(1)}\text{nonst}}\bigl{|}_{x\to\infty}&=-2C_{2}\xi^{(\mu}\omega^{\nu)}+\frac{C_{2}}{5}\delta^{(\mu}_{k}\varepsilon^{\nu)0lm}(-ik_{l})h^{k}_{m}\,.\\ \end{split} (114)

These results, combined with the static parts (74), yield the O()O(\hbar) contributions of Eq. (110), as follows:

J(1)dynμ=0,T(1)dyn0μ=0.\begin{split}J^{\mu}_{{(1)}\text{dyn}}=0\,,\quad T^{0\mu}_{{(1)}\text{dyn}}=0\,.\end{split} (115)

Therefore we conclude that the CVE vanishes in the dynamical limit. This is consistent with the diagrammatic calculation in Ref. Landsteiner:2013aba .

At O(2)O(\hbar^{2}), from the Wigner function (95), the nonstatic parts are evaluated as

J(2)nonstμ=C04[12Rμν(ξνI+δkνIk)14R(ξμI+δkμIk)+Rα0(ξμξαI+(ξμδkα+ξαδkμ)Ik+δjμδkαIjk)+Rj0k0(ξμIjk+δiμIijk)],T(2)nonstμν=(2C1)[116R(ξμξνI+2ξ(μδkν)Ik+δjμδkνIjk)+38R0α(ξμξνξαI+(2ξ(μδkν)ξα+ξμξνδkα)Ik+(2δj(μξν)δkα+δjμδkνξα)Ijk+δiμδjνδkαIijk)+12Rk0l0(ξμξνIkl+2ξ(μδjν)Ijkl+δiμδjνIijkl)].\begin{split}J^{\mu}_{{(2)}\text{nonst}}&=-\frac{C_{0}}{4}\biggl{[}\frac{1}{2}{R^{\mu}}_{\nu}\biggl{(}\xi^{\nu}I+\delta^{\nu}_{k}I^{k}\biggr{)}-\frac{1}{4}R\biggl{(}\xi^{\mu}I+\delta^{\mu}_{k}I^{k}\biggr{)}\\ &\qquad+R_{\alpha 0}\biggl{(}\xi^{\mu}\xi^{\alpha}I+(\xi^{\mu}\delta^{\alpha}_{k}+\xi^{\alpha}\delta^{\mu}_{k})I^{k}+\delta^{\mu}_{j}\delta^{\alpha}_{k}I^{jk}\biggr{)}+R_{j0k0}\biggl{(}\xi^{\mu}I^{jk}+\delta^{\mu}_{i}I^{ijk}\biggr{)}\biggr{]}\,,\\ T^{\mu\nu}_{{(2)}\text{nonst}}&=(-2C_{1})\biggl{[}-\frac{1}{16}R\biggl{(}\xi^{\mu}\xi^{\nu}I+2\xi^{(\mu}\delta^{\nu)}_{k}I^{k}+\delta^{\mu}_{j}\delta^{\nu}_{k}I^{jk}\biggr{)}\\ &\quad+\frac{3}{8}R_{0\alpha}\biggl{(}\xi^{\mu}\xi^{\nu}\xi^{\alpha}I+(2\xi^{(\mu}\delta^{\nu)}_{k}\xi^{\alpha}+\xi^{\mu}\xi^{\nu}\delta^{\alpha}_{k})I^{k}+(2\delta^{(\mu}_{j}\xi^{\nu)}\delta^{\alpha}_{k}+\delta^{\mu}_{j}\delta^{\nu}_{k}\xi^{\alpha})I^{jk}\\ &\qquad+\delta^{\mu}_{i}\delta^{\nu}_{j}\delta^{\alpha}_{k}I^{ijk}\biggr{)}+\frac{1}{2}R_{k0l0}\biggl{(}\xi^{\mu}\xi^{\nu}I^{kl}+2\xi^{(\mu}\delta^{\nu)}_{j}I^{jkl}+\delta^{\mu}_{i}\delta^{\nu}_{j}I^{ijkl}\biggr{)}\biggr{]}\,.\end{split} (116)

In the dynamical limit xx\to\infty, we obtain

J(2)nonstμ|x=C0[16ξμR00+160(2R0μ+ξμR)],T(2)nonstμν|x=C1[ξμξν(13140R2435R00)+730ξ(μR0ν)ημν(11420R16105R00)215Rμ0ν0].\begin{split}J^{\mu}_{{(2)}\text{nonst}}\bigl{|}_{x\to\infty}&=C_{0}\biggl{[}-\frac{1}{6}\xi^{\mu}R_{00}+\frac{1}{60}\biggl{(}-2{R_{0}}^{\mu}+\xi^{\mu}R\biggr{)}\biggr{]}\,,\\ T^{\mu\nu}_{{(2)}\text{nonst}}\bigl{|}_{x\to\infty}&=C_{1}\biggl{[}\xi^{\mu}\xi^{\nu}\biggl{(}\frac{13}{140}R-\frac{24}{35}R_{00}\biggr{)}+\frac{7}{30}\xi^{(\mu}{R_{0}}^{\nu)}\\ &\qquad\qquad-\eta^{\mu\nu}\biggl{(}\frac{11}{420}R-\frac{16}{105}R_{00}\biggr{)}-\frac{2}{15}R^{\mu 0\nu 0}\biggr{]}\,.\end{split} (117)

Here we used

1xR0j0kk^k=kjk0R00Rj0,1x2R0j0kk^jk^k=(11x2)R0012R,R=2R00+2kkk0R0k,\frac{1}{x}R_{0j0k}{\hat{k}}^{k}=\frac{k_{j}}{k_{0}}R_{00}-R_{j0}\,,\quad\frac{1}{x^{2}}R_{0j0k}{\hat{k}}^{j}{\hat{k}}^{k}=\biggl{(}1-\frac{1}{x^{2}}\biggr{)}R_{00}-\frac{1}{2}R\,,\quad R=2R_{00}+2\frac{k_{k}}{k_{0}}{R_{0}}^{k}\,, (118)

which follow from the second Bianchi identity. Combining these with the static contribution (78), we write the O(2)O(\hbar^{2}) contributions of Eq. (110) as

J(2)dynμ=C020[Rμαξα12ξμR],T(2)dynμν=C1[112Rμν+1105Rξμξν+13840Rημν+115Rα(μξν)ξα2105Rαβξαξβξμξν170Rαβξαξβημν+130Rμανβξαξβ].\begin{split}J^{\mu}_{{(2)}\text{dyn}}&=\frac{C_{0}}{20}\biggl{[}{R^{\mu}}_{\alpha}\xi^{\alpha}-\frac{1}{2}\xi^{\mu}R\biggr{]}\,,\\ T^{\mu\nu}_{{(2)}\text{dyn}}&=C_{1}\biggl{[}-\frac{1}{12}R^{\mu\nu}+\frac{1}{105}R\xi^{\mu}\xi^{\nu}+\frac{13}{840}R\eta^{\mu\nu}+\frac{1}{15}R^{\alpha(\mu}\xi^{\nu)}\xi_{\alpha}\\ &\qquad\quad-\frac{2}{105}R^{\alpha\beta}\xi_{\alpha}\xi_{\beta}\xi^{\mu}\xi^{\nu}-\frac{1}{70}R^{\alpha\beta}\xi_{\alpha}\xi_{\beta}\eta^{\mu\nu}+\frac{1}{30}R^{\mu\alpha\nu\beta}\xi_{\alpha}\xi_{\beta}\biggr{]}\,.\end{split} (119)

We note that Eqs. (101)-(104) and Eqs. (106)-(108) again result in the conservation laws μJ(1)μ=μJ(2)μ=0\nabla_{\mu}J^{\mu}_{{(1)}}=\nabla_{\mu}J^{\mu}_{{(2)}}=0, μT(1)μν=μT(2)μν=0\nabla_{\mu}T^{\mu\nu}_{{(1)}}=\nabla_{\mu}T^{\mu\nu}_{{(2)}}=0 and Tμμ(1)=Tμμ(2)=0{T^{\mu}}_{\mu{(1)}}={T^{\mu}}_{\mu{(2)}}=0 for arbitrary kμk^{\mu}. In the next section, we discuss some implications of (119) in the fluid frame [see Eq. (7)].

7 Fluid frame

The fermionic system under a background fluid is pedagogical and informative to show the novelty of the gravity-induced transport phenomena given by Eqs. (78) and (119). In general, the effect of the fluid can be translated to that of an effective curved geometry with the following metric Banerjee:2012iz :

g00=1+h00(t,𝒙),g0i=h0i(t,𝒙),gij=ηij.g_{00}=1+h_{00}(t,{\boldsymbol{x}})\,,\quad g_{0i}=h_{0i}(t,{\boldsymbol{x}})\,,\quad g_{ij}=\eta_{ij}\,. (120)

Adopting this metric, we represent the temperature gradient Luttinger:1964zz and the fluid vorticity as

iT/T¯=12ig00,ωi=12ε0ijkjgk0.\partial_{i}T/\bar{T}=-\frac{1}{2}\partial_{i}g_{00}\,,\quad\omega^{i}=-\frac{1}{2}\varepsilon^{0ijk}\partial_{j}g_{k0}\,. (121)

with T¯\bar{T} being the global temperature. Alternatively, the present coordinate describes the system under the gravitoelectromagnetic fields i=12ig00\mathcal{E}^{i}=-\frac{1}{2}\partial^{i}g_{00} and i=12εijkjgk0\mathcal{B}^{i}=-\frac{1}{2}\varepsilon^{ijk}\partial_{j}g_{k0}. The nonvanishing components of the curvature tensors read

Ri0j0=Rij=ijTT¯0ϵij,R00=12R=2TT¯0ϵjj,R0i=(×𝝎)i,R_{i0j0}=R_{ij}=-\frac{\partial_{i}\partial_{j}T}{\bar{T}}-\partial_{0}\epsilon_{ij}\,,\quad R_{00}=\frac{1}{2}R=\frac{{\boldsymbol{\nabla}}^{2}T}{\bar{T}}-\partial_{0}{\epsilon_{j}}^{j}\,,\quad R_{0i}=({\boldsymbol{\nabla}}\times{\boldsymbol{\omega}})_{i}\,, (122)

with ϵij=12(ih0j+jh0i)\epsilon_{ij}=\frac{1}{2}(\partial_{i}h_{0j}+\partial_{j}h_{0i}).

Inserting Eq. (122) into Eqs. (78) and (119), we readily obtain the transport phenomena under the temperature gradient and inhomogeneous vorticity. In the static limit (or equivalently, for the stationary metric 0hμν=0\partial_{0}h_{\mu\nu}=0), Eq. (78) is written explicitly as

J(2)eq0=C062TT¯,J(2)eqi=C012(×𝝎)i,\displaystyle\displaystyle J^{0}_{{(2)}\text{eq}}=\frac{C_{0}}{6}\frac{{\boldsymbol{\nabla}}^{2}T}{\bar{T}}\,,\quad J^{i}_{{(2)}\text{eq}}=\frac{C_{0}}{12}({\boldsymbol{\nabla}}\times{\boldsymbol{\omega}})^{i}\,,
T(2)eq00=C162TT¯,T(2)eq0i=C16(×𝝎)i,T(2)eqij=C112T¯(ij+ηij2)T,\displaystyle\displaystyle T^{00}_{{(2)}\text{eq}}=\frac{C_{1}}{6}\frac{{\boldsymbol{\nabla}}^{2}T}{\bar{T}}\,,\quad T^{0i}_{{(2)}\text{eq}}=-\frac{C_{1}}{6}({\boldsymbol{\nabla}}\times{\boldsymbol{\omega}})^{i}\,,\quad T^{ij}_{{(2)}\text{eq}}=-\frac{C_{1}}{12\bar{T}}(\partial^{i}\partial^{j}+\eta^{ij}{\boldsymbol{\nabla}}^{2})T\,, (123)

with C0=μ/2π2C_{0}=\mu/2\pi^{2} and C1=μ2/4π2+T2/12C_{1}=\mu^{2}/4\pi^{2}+T^{2}/12. Similarly, from the expression in the dynamical limit (119), we find

J(2)dyn0=0,J(2)dyni=C020(×𝝎)i,\displaystyle\displaystyle J^{0}_{{(2)}\text{dyn}}=0\,,\quad J^{i}_{{(2)}\text{dyn}}=\frac{C_{0}}{20}({\boldsymbol{\nabla}}\times{\boldsymbol{\omega}})^{i}\,,
T(2)dyn00=0,/T(2)dyn0i=C120(×𝝎)i,T(2)dynij=C120[1T¯(ij+13ηij2)T+0σij],\displaystyle\displaystyle T^{00}_{{(2)}\text{dyn}}=0\,,\quad/T^{0i}_{{(2)}\text{dyn}}=-\frac{C_{1}}{20}({\boldsymbol{\nabla}}\times{\boldsymbol{\omega}})^{i}\,,\quad T^{ij}_{{(2)}\text{dyn}}=\frac{C_{1}}{20}\biggl{[}\frac{1}{\bar{T}}\biggl{(}\partial^{i}\partial^{j}+\frac{1}{3}\eta^{ij}{\boldsymbol{\nabla}}^{2}\biggr{)}T+\partial_{0}\sigma^{ij}\biggr{]}\,,

where we introduce the shear tensor:

σij=ϵij13ηijϵkk.\sigma^{ij}=\epsilon^{ij}-\frac{1}{3}\eta^{ij}{\epsilon_{k}}^{k}\,. (125)

The corresponding vector and axial-vector currents are obtained by replacing C0C_{0} with C0,V/A=μV/A/π2C_{0,V/A}=\mu_{V/A}/\pi^{2}, and C1C_{1} with C1,V=(μV2+μA2)/2π2+T2/6C_{1,V}=(\mu_{V}^{2}+\mu^{2}_{A})/2\pi^{2}+T^{2}/6 and C1,A=μVμA/π2C_{1,A}=\mu_{V}\mu_{A}/\pi^{2} respectively, as we have done to get Eq. (79). Several comments are in order.

We again emphasize that Eq. (7) represents equilibrium transport phenomena, like the CME and CVE. It is also intriguing to notice the difference between these and Fourier’s law. For the former, the currents in Eq. (7) come from the vorticity, namely, the magnetic part \mathcal{B} of the gravity. This background source supplies no energy to particles, and thus the currents become finite even in equilibrium. On the other hand, the latter is the current generation by the temperature gradient. This electric part \mathcal{E} of the gravitational field gives an energy to particles. Therefore the Fourier’s law is dissipative and prohibited in equilibrium.

For an static and spatially inhomogeneous vorticity 𝝎(𝒙){\boldsymbol{\omega}}({\boldsymbol{x}}), there emerges the nonvanishing charge current J(2)eqiJ^{i}_{{(2)}\text{eq}} and the energy current T(2)eq0iT^{0i}_{{(2)}\text{eq}}, on top of the contributions from the CVE (74). Unlike the vector part of the CVE, the curvature-induced currents (79) or (7) does not require μA0\mu_{A}\neq 0. In the system without the chiral imbalance, hence, J(2)eqiJ^{i}_{{(2)}\text{eq}} and T(2)eq0iT^{0i}_{{(2)}\text{eq}} are the leading vortical contributions. In the dynamical limit, such second-order contributions become more important, since the CVE is washed out as shown in Eq. (115) while the currents in Eq. (7) are not.

Under the correspondence between magnetic field and vorticity, one would think that the charge current J(2)eq/dyniJ^{i}_{{(2)}\text{eq}/\text{dyn}} is the gravitational analogue of Ampère’s law: ×𝓑=×𝝎=𝑱{\boldsymbol{\nabla}}\times\boldsymbol{\mathcal{B}}={\boldsymbol{\nabla}}\times{\boldsymbol{\omega}}={\boldsymbol{J}}. The situation is however not so trivial since J(2)eq/dyniJ^{i}_{{(2)}\text{eq}/\text{dyn}} is opposite-signed against the energy current T(2)eq/dyn0iT^{0i}_{{(2)}\text{eq}/\text{dyn}} (for μ>0\mu>0). Namely, Eqs. (7) and (7) cannot be explained based on the naive picture that a particle’s momentum carries both charge and energy. This curious flow dynamics essentially comes from the quantum effects through the spin-curvature coupling. We should emphasize that such an antiparallel charge-energy flow is not restricted in the present coordinate, but more generally admitted in a lot of curved spacetime; this phenomenon always takes place as long as R0i0{R_{0}}^{i}\neq 0, as shown in Eqs. (78) and (119).

It is worthwhile to mention the feedback to the gravitational field from Eq. (78). In our sign convention, the Einstein field equation is given by Rμν12gμνR=8πGTμνR_{\mu\nu}-\frac{1}{2}g_{\mu\nu}R=-8\pi GT_{\mu\nu} with the gravitational constant GG birrell1984quantum . Following this, the induced Ricci tensor reads R0iindαR0iR^{\text{ind}}_{0i}\sim\alpha R_{0i} with a positive coefficient α>0\alpha>0. Hence, the initial gravitational field is enhanced, which evokes the possibility of instability. We will revisit and analyze more precisely the above brief argument in the future, including the existence of a novel collective dynamics Akamatsu:2013pjd in a gravitational plasma Nachbagauer:1995wn ; deAlmeida:1993wy .

One might think that Eq. (7) is unrelated to an anomaly. Indeed, Eq. (7) would be irrelevant to the chiral anomaly, according to the analysis of discrete symmetry Kharzeev:2011ds . Nevertheless, this fact is not sufficient to conclude the irrelevance to anomaly at all, as for the temperature dependent part of the CVE Golkar:2012kb ; Golkar:2015oxw ; Chowdhury:2016cmh ; Glorioso:2017lcn . We also mention that the transport coefficients C0C_{0} and C1C_{1} are time-reversal even quantities, which could be associated with their nondissipative nature similarly to those of the CME and CVE Kharzeev:2011ds . It should be required to clarify the anomalous aspect of Eq. (7) from different approaches, such as hydrodynamics. In the sense that they are of the higher-order of the derivative counting, usual hydrodynamics can neglect Eqs. (7) and (7). This would not be the case, however, if these phenomena originate from quantum anomaly like the CME and CVE.

These novel contributions (7) lead to several implications in relativistic many-body systems where an inhomogeneous fluid vorticity is experimentally generated. In rotating quark-gluon plasma, there emerges the quadrupole configuration of the vorticity along the beam direction Adam:2019srw ; Becattini:2017gcx ; Wei:2018zfb ; Fu:2020oxj . Thus, on the transverse plane to the beam direction, the inhomogeneous vorticities generate the charge current JJ_{\perp} and the energy current JϵJ^{\epsilon}_{\perp}, as depicted in Fig. 1. As a brief argument, we may estimate the scale of the vorticity gradient to be the inverse of the hot matter size. Indeed, at the collision energy s=19.6GeV\sqrt{s}=19.6\,\text{GeV}, the gradient of the vorticity is estimated to be (×𝝎)/𝝎0.2fm0.5fm40MeV100MeV({\boldsymbol{\nabla}}\times{\boldsymbol{\omega}})/{\boldsymbol{\omega}}\approx 0.2\,\text{fm}-0.5\,\text{fm}\approx 40\,\text{MeV}-100\,\text{MeV} Wei:2018zfb . Although the whole magnitudes of JJ_{\perp} and JϵJ^{\epsilon}_{\perp} are dependent on the scale of 𝝎{\boldsymbol{\omega}}, hence, these are nonnegligible compared with the CVE.

Refer to caption
Figure 1: Flow directions of JJ_{\perp} (left) and JϵJ^{\epsilon}_{\perp} (right). The horizontal axis corresponds to the reaction plane of heavy-ion collisions. The quadrupole vorticity structure is based on the measurement by STAR collaboration Adam:2019srw

On top of the charge and energy currents, the stress tensor T(2)eqijT^{ij}_{{(2)}\text{eq}} is also induced. Let us consider a cylindrical system along the zz direction with a spatially inhomogeneous temperature T(z)T(z). From the vector part of the energy-momentum tensor in Eq. (7), the temperature gradient yields the correction to the transverse pressure P(z)=C1,V12T′′(z)/T¯P_{\perp}(z)=\frac{C_{1,V}}{12}T^{\prime\prime}(z)/\bar{T}. When the temperature takes a Gaussian form T(z)=T¯ez2/2σ2T(z)=\bar{T}e^{-z^{2}/2\sigma^{2}}, we get P(z)=C1,V12ez2/2σ2(z2σ2)/(3σ2)P_{\perp}(z)=\frac{C_{1,V}}{12}e^{-z^{2}/2\sigma^{2}}(z^{2}-\sigma^{2})/(3\sigma^{2}), which has the minima P(0)=C1,V12σ2<0P_{\perp}(0)=-\frac{C_{1,V}}{12}\sigma^{-2}<0 and maxima P(σ)=C1,V12e3/2σ2>0P_{\perp}(\sigma)=\frac{C_{1,V}}{12}e^{-3/2}\\ \sigma^{-2}>0. Such a pressure correction is detectable in Weyl/Dirac semimetal experiments, similarly to the usual thermoelectric transport phenomena Lundgren:2014hra ; liang2017anomalous .

In table-top experiments, an inhomogeneous and dynamical vorticity can be generated by an acoustic surface wave. We consider a transverse wave propagating on the xyxy surface 9780750626330 ; PhysRevLett.119.077202 ; PhysRevB.87.180402 of Weyl/Dirac semimetals. Also we prepare the wave propagating along the xx direction, and its amplitude is normal to the surface, i.e., its displacement vector is given by 𝒖=(0,0,u){\boldsymbol{u}}=(0,0,u) with u=u¯eik0t+ikxκzu=\bar{u}e^{-ik_{0}t+ikx-\kappa z}. Here eκze^{-\kappa z} reflects unpenetrating into the material. Now the response to this surface wave can be evaluated in the coordinate space described by gμν=ημν+hμνg_{\mu\nu}=\eta_{\mu\nu}+h_{\mu\nu} with h0z=u˙=ik0uh_{0z}=-\dot{u}=ik_{0}u, hxz=xu=ikuh_{xz}=-\partial_{x}{u}=-iku, hzz=zu=κuh_{zz}=-\partial_{z}{u}=\kappa u and other components of hμνh_{\mu\nu} vanishing. From Eq. (7) together with the Wick rotation 3κ\partial_{3}\to\kappa, we get the charge and energy currents: J(2)dynx=C02012k0kκuJ^{x}_{{(2)}\text{dyn}}=\frac{C_{0}}{20}\frac{1}{2}k_{0}k\kappa u, J(2)dynz=C02012ik0k2uJ^{z}_{{(2)}\text{dyn}}=\frac{C_{0}}{20}\frac{1}{2}ik_{0}k^{2}u and T(2)dyn0x=C04012k0kκuT^{0x}_{{(2)}\text{dyn}}=-\frac{C_{0}}{40}\frac{1}{2}k_{0}k\kappa u, T(2)dyn0z=C04012ik0k2uT^{0z}_{{(2)}\text{dyn}}=-\frac{C_{0}}{40}\frac{1}{2}ik_{0}k^{2}u. The flows normal to the fluid velocity u˙\dot{u} are induced by the gravitational curvature via quantum effects. We note that the flows parallel to the fluid velocity are induced from classical contributions.

Acknowledgements.
This work was supported by JSPS KAKENHI Grant Numbers 17H06462 and 18H01211. K. M. was supported by Special Postdoctoral Researcher (SPDR) Program of RIKEN.

Appendix A Equilibrium Wigner function (66)

In this appendix, we show the concrete expression of (2)μ\mathcal{R}^{\mu}_{(2)} at equilibrium defined by Eqs. (59), (62), (65) and (71). We decompose (2)μ\mathcal{R}^{\mu}_{(2)} in Eq. (42) into the frame-(in)dependent and the ϕ(2)\phi_{(2)} part:

(2)μ=(2)indepμ+(2)depμ+2πδ(p2)pμϕ(2).\mathcal{R}^{\mu}_{(2)}=\mathcal{R}^{\mu}_{{(2)}\text{indep}}+\mathcal{R}^{\mu}_{{(2)}\text{dep}}+2\pi\delta(p^{2})p^{\mu}\phi_{(2)}\,. (126)

The first term reads

(2)indepμ=2πp2[pμQp+2pν(T[μpν]+Sαμνpα)]δ(p2)f(0)=2π24[5RαμαpRαβγμpαβpγppμp2(2R+6Rαβpαβp+2Rαβγδpαpγpβpδ)6p2Rαβγμpαpγβp]δ(p2)f(0)=2πδ(p2)[f(0)(12p2Rμαpα112p2Rpμ+23(p2)2Rαβpμpαpβ)+f(0)(524Rμαβα16p2Rαβγμpαββpγ14p2Rαβpμpαββ)+f(0)′′(124Rαβγμpαβββγ112p2Rαβγδpμpαpγβββδ)].\begin{split}\mathcal{R}^{\mu}_{{(2)}\text{indep}}&=\frac{2\pi}{p^{2}}\biggl{[}-p^{\mu}Q\cdot p+2p_{\nu}\Bigl{(}T^{[\mu}p^{\nu]}+S^{\alpha\mu\nu}p_{\alpha}\Bigr{)}\biggr{]}\delta(p^{2})f_{(0)}\\ &=\frac{2\pi}{24}\biggl{[}5R^{\alpha\mu}\partial^{p}_{\alpha}-R^{\alpha\beta\gamma\mu}p_{\alpha}\partial^{p}_{\beta}\partial_{\gamma}^{p}\ -\frac{p^{\mu}}{p^{2}}\bigl{(}2R+6R^{\alpha\beta}p_{\alpha}\partial_{\beta}^{p}+2R_{\alpha\beta\gamma\delta}p^{\alpha}p^{\gamma}\partial_{p}^{\beta}\partial^{\delta}_{p}\bigr{)}\\ &\qquad-\frac{6}{p^{2}}R^{\alpha\beta\gamma\mu}p_{\alpha}p_{\gamma}\partial^{p}_{\beta}\biggr{]}\delta(p^{2})f_{(0)}\\ &=2\pi\delta(p^{2})\Biggl{[}f_{(0)}\biggl{(}-\frac{1}{2p^{2}}R^{\mu\alpha}p_{\alpha}-\frac{1}{12p^{2}}R\,p^{\mu}+\frac{2}{3(p^{2})^{2}}R^{\alpha\beta}p^{\mu}p_{\alpha}p_{\beta}\biggr{)}\\ &\qquad\qquad\quad+f_{(0)}^{\prime}\biggl{(}\frac{5}{24}R^{\mu\alpha}\beta_{\alpha}-\frac{1}{6p^{2}}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}p_{\gamma}-\frac{1}{4p^{2}}R^{\alpha\beta}p^{\mu}p_{\alpha}\beta_{\beta}\biggr{)}\\ &\qquad\qquad\quad+f_{(0)}^{\prime\prime}\biggl{(}-\frac{1}{24}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}\beta_{\gamma}-\frac{1}{12p^{2}}R_{\alpha\beta\gamma\delta}p^{\mu}p^{\alpha}p^{\gamma}\beta^{\beta}\beta^{\delta}\biggr{)}\Biggr{]}\,.\end{split} (127)

The frame-dependent part is further decomposed as

(2)depμ=2πδ(p2)(r1μ+r2μ+r3μ),\begin{split}\mathcal{R}^{\mu}_{{(2)}\text{dep}}&=2\pi\delta(p^{2})\Bigl{(}r^{\mu}_{1}+r^{\mu}_{2}+r^{\mu}_{3}\Bigr{)}\,,\end{split} (128)

where we define

r1μ:=pμΣνρuDν(f(0)ερσλη4pnnσλβη)+12ΣuμνDν(f(0)Σnρσρβσ),r2μ:=12p2εμνρσpνDρΣσλnDλf(0),r3μ:=1p2Σuμν(12R~αβνρpρpαpβ+pDΣνρnDρ)f(0).\begin{split}r^{\mu}_{1}&:=p^{\mu}\Sigma^{u}_{\nu\rho}D^{\nu}\biggl{(}f^{\prime}_{(0)}\frac{\varepsilon^{\rho\sigma\lambda\eta}}{4\,p\cdot n}n_{\sigma}\nabla_{\lambda}\beta_{\eta}\biggr{)}+\frac{1}{2}\Sigma_{u}^{\mu\nu}D_{\nu}\Bigl{(}f^{\prime}_{(0)}\Sigma_{n}^{\rho\sigma}\nabla_{\rho}\beta_{\sigma}\Bigr{)}\,,\\ r^{\mu}_{2}&:=\frac{1}{2p^{2}}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}D_{\rho}\Sigma^{n}_{\sigma\lambda}D^{\lambda}f_{(0)}\,,\\ r^{\mu}_{3}&:=-\frac{1}{p^{2}}\Sigma^{\mu\nu}_{u}\biggl{(}\frac{1}{2}{\tilde{R}}_{\alpha\beta\nu\rho}p^{\rho}p^{\alpha}\partial^{\beta}_{p}+p\cdot D\Sigma^{n}_{\nu\rho}D^{\rho}\biggr{)}f_{(0)}\,.\end{split} (129)

For the equilibrium distribution f(0)f_{(0)} in Eq. (59), we reduce r2μr^{\mu}_{2} and r3μr^{\mu}_{3} to

r2μ=12p2εμνρσpνDρΣσλnpηλβηf(0)=18p2εμνρσpνDρεληστpτλβηf(0)+18εμνρσpνDρεληστnτpnλβηf(0),\begin{split}r^{\mu}_{2}&=\frac{1}{2p^{2}}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}D_{\rho}\Sigma^{n}_{\sigma\lambda}p_{\eta}\nabla^{\lambda}\beta^{\eta}f^{\prime}_{(0)}\\ &=-\frac{1}{8p^{2}}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}D_{\rho}\varepsilon_{\lambda\eta\sigma\tau}p^{\tau}\nabla^{\lambda}\beta^{\eta}f^{\prime}_{(0)}+\frac{1}{8}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}D_{\rho}\varepsilon_{\lambda\eta\sigma\tau}\frac{n^{\tau}}{p\cdot n}\nabla^{\lambda}\beta^{\eta}f^{\prime}_{(0)}\,,\end{split} (130)

and

r3μ=1p2Σuμν[12R~αβνρpρpαββf(0)+pDΣνρnpλρβλf(0)]=14f(0)ΣuμνερλντpDnτpnρβλ=14(2Σuμ[νpσ]+Σuμσpν)ερλντDσf(0)nτpnρβλ=r1μ+18ενσμηpηερλντDσf(0)nτpnρβλ,\begin{split}r^{\mu}_{3}&=-\frac{1}{p^{2}}\Sigma_{u}^{\mu\nu}\biggl{[}\frac{1}{2}\tilde{R}_{\alpha\beta\nu\rho}p^{\rho}p^{\alpha}\beta^{\beta}f^{\prime}_{(0)}+p\cdot D\Sigma^{n}_{\nu\rho}p_{\lambda}\nabla^{\rho}\beta^{\lambda}f^{\prime}_{(0)}\biggr{]}\\ &=-\frac{1}{4}f^{\prime}_{(0)}\Sigma_{u}^{\mu\nu}\varepsilon_{\rho\lambda\nu\tau}p\cdot D\frac{n^{\tau}}{p\cdot n}\nabla^{\rho}\beta^{\lambda}\\ &=-\frac{1}{4}(2\Sigma_{u}^{\mu[\nu}p^{\sigma]}+\Sigma_{u}^{\mu\sigma}p^{\nu})\varepsilon_{\rho\lambda\nu\tau}D_{\sigma}f^{\prime}_{(0)}\frac{n^{\tau}}{p\cdot n}\nabla^{\rho}\beta^{\lambda}\\ &=-r^{\mu}_{1}+\frac{1}{8}\varepsilon^{\nu\sigma\mu\eta}p_{\eta}\varepsilon_{\rho\lambda\nu\tau}D_{\sigma}f^{\prime}_{(0)}\frac{n^{\tau}}{p\cdot n}\nabla^{\rho}\beta^{\lambda}\,,\end{split} (131)

where the p2p^{2} term is dropped, and we utilize μνβρ=βλRλμνρ\nabla_{\mu}\nabla_{\nu}\beta_{\rho}=-\beta^{\lambda}R_{\lambda\mu\nu\rho} and Eq. (45). The frame-dependent part hence becomes

(2)depμ=2πδ(p2)8p2εμνρσpνDρεληστpτλβηf(0)=2πδ(p2)[f(0)(14p2pαββpγRαβγμ14Rμνβν+pμ4p2Rαβpαββ)+f(0)′′(14ρβμpνρβν+pμ4p2pνρβνpσρβσ)].\begin{split}\mathcal{R}^{\mu}_{{(2)}\text{dep}}&=-\frac{2\pi\delta(p^{2})}{8p^{2}}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}D_{\rho}\varepsilon_{\lambda\eta\sigma\tau}p^{\tau}\nabla^{\lambda}\beta^{\eta}f^{\prime}_{(0)}\\ &=2\pi\delta(p^{2})\biggl{[}f_{(0)}^{\prime}\biggl{(}\frac{1}{4p^{2}}p_{\alpha}\beta_{\beta}p_{\gamma}R^{\alpha\beta\gamma\mu}-\frac{1}{4}R^{\mu\nu}\beta_{\nu}+\frac{p^{\mu}}{4p^{2}}R^{\alpha\beta}p_{\alpha}\beta_{\beta}\biggr{)}\\ &\quad+f^{\prime\prime}_{(0)}\biggl{(}-\frac{1}{4}\nabla^{\rho}\beta^{\mu}p^{\nu}\nabla_{\rho}\beta_{\nu}+\frac{p^{\mu}}{4p^{2}}p_{\nu}\nabla^{\rho}\beta^{\nu}p^{\sigma}\nabla_{\rho}\beta_{\sigma}\biggr{)}\biggr{]}\,.\end{split} (132)

In the above equation, the frame dependence totally vanishes, as it should. Eventually, (2)μ\mathcal{R}^{\mu}_{(2)} is written as Eq. (66).

Appendix B Equilibrium kinetic equation (67)

In this appendix, we derive the kinetic equation (67). In later use, we recall the second Bianchi identity for the Riemann tensor:

αRμνβγ+βRμνγα+γRμναβ=0,\nabla_{\alpha}R_{\mu\nu\beta\gamma}+\nabla_{\beta}R_{\mu\nu\gamma\alpha}+\nabla_{\gamma}R_{\mu\nu\alpha\beta}=0\,, (133)

which implies

μRμν=12νR,μRρσμν=ρRσνσRρν.\nabla_{\mu}R^{\mu\nu}=\frac{1}{2}\nabla^{\nu}R\,,\quad\nabla_{\mu}R^{\rho\sigma\mu\nu}=\nabla^{\rho}R^{\sigma\nu}-\nabla^{\sigma}R^{\rho\nu}\,. (134)

Using μβν=νβμ\nabla_{\mu}\beta_{\nu}=-\nabla_{\nu}\beta_{\mu}, Rα[μν]β=12RαβμνR_{\alpha[\mu\nu]\beta}=-\frac{1}{2}R_{\alpha\beta\mu\nu} and Eq. (134), we evaluate each term in the kinetic equation (8) as follows:

18λRμνpλpνpμf(0)δ(p2)=f(0)(12δ(p2)pR12δ′′(p2)pRαβpαpβ)+f(0)(316δ(p2)βR14δ(p2)pRαβpαββ14δ(p2)βRαβpαpβ)+f(0)′′(18δ(p2)βRαβpαββ),\begin{split}&-\frac{1}{8}\nabla_{\lambda}R_{\mu\nu}\partial_{p}^{\lambda}\partial_{p}^{\nu}p^{\mu}f_{(0)}\delta(p^{2})\\ &=f_{(0)}\biggl{(}-\frac{1}{2}\delta^{\prime}(p^{2})p\cdot\nabla R-\frac{1}{2}\delta^{\prime\prime}(p^{2})p\cdot\nabla R^{\alpha\beta}p_{\alpha}p_{\beta}\biggr{)}\\ &\quad+f_{(0)}^{\prime}\biggl{(}-\frac{3}{16}\delta(p^{2})\beta\cdot\nabla R-\frac{1}{4}\delta^{\prime}(p^{2})p\cdot\nabla R^{\alpha\beta}p_{\alpha}\beta_{\beta}-\frac{1}{4}\delta^{\prime}(p^{2})\beta\cdot\nabla R^{\alpha\beta}p_{\alpha}p_{\beta}\biggr{)}\\ &\quad+f_{(0)}^{\prime\prime}\biggl{(}-\frac{1}{8}\delta(p^{2})\beta\cdot\nabla R^{\alpha\beta}p_{\alpha}\beta_{\beta}\biggr{)}\,,\end{split} (135)
124λRρσμνpλpνpσpρpμf(0)δ(p2)=f(0)(16δ(p2)pR+16δ′′(p2)pRαβpαpβ)+f(0)(112δ(p2)βR+16δ(p2)pRαβpαββ+112δ(p2)βRαβpαpβ)+f(0)′′(16δ(p2)βRαβpαββ112δ(p2)pRαββαββ112δ(p2)pRρσμνpμβνpρβσ)+f(0)′′′(124δ(p2)βRρσμνpμβνpρβσ),\begin{split}&-\frac{1}{24}\nabla_{\lambda}R_{\rho\sigma\mu\nu}\partial_{p}^{\lambda}\partial_{p}^{\nu}\partial_{p}^{\sigma}p^{\rho}p^{\mu}f_{(0)}\delta(p^{2})\\ &=f_{(0)}\biggl{(}\frac{1}{6}\delta^{\prime}(p^{2})p\cdot\nabla R+\frac{1}{6}\delta^{\prime\prime}(p^{2})p\cdot\nabla R^{\alpha\beta}p_{\alpha}p_{\beta}\biggr{)}\\ &\quad+f_{(0)}^{\prime}\biggl{(}\frac{1}{12}\delta(p^{2})\beta\cdot\nabla R+\frac{1}{6}\delta^{\prime}(p^{2})p\cdot\nabla R^{\alpha\beta}p_{\alpha}\beta_{\beta}+\frac{1}{12}\delta^{\prime}(p^{2})\beta\cdot\nabla R^{\alpha\beta}p_{\alpha}p_{\beta}\biggr{)}\\ &\quad+f_{(0)}^{\prime\prime}\biggl{(}\frac{1}{6}\delta(p^{2})\beta\cdot\nabla R^{\alpha\beta}p_{\alpha}\beta_{\beta}-\frac{1}{12}\delta(p^{2})p\cdot\nabla R_{\alpha\beta}\beta^{\alpha}\beta^{\beta}-\frac{1}{12}\delta^{\prime}(p^{2})p\cdot\nabla R_{\rho\sigma\mu\nu}p^{\mu}\beta^{\nu}p^{\rho}\beta^{\sigma}\biggr{)}\\ &\quad+f_{(0)}^{\prime\prime\prime}\biggl{(}-\frac{1}{24}\delta(p^{2})\beta\cdot\nabla R_{\rho\sigma\mu\nu}p^{\mu}\beta^{\nu}p^{\rho}\beta^{\sigma}\biggr{)}\,,\end{split} (136)
18RρσμνpνpσDρpμf(0)δ(p2)=f(0)′′(316δ(p2)Rρσμνpμβνρβσ)+(18δ(p2)Rαρβα+14δ(p2)Rρσμνpμβνpσ)Dρf(0)+(18δ(p2)Rρσμνpμβνβσ)Dρf(0)′′,\begin{split}&\frac{1}{8}{R^{\rho}}_{\sigma\mu\nu}\partial_{p}^{\nu}\partial_{p}^{\sigma}D_{\rho}p^{\mu}f_{(0)}\delta(p^{2})\\ &=f_{(0)}^{\prime\prime}\biggl{(}\frac{3}{16}\delta(p^{2})R^{\rho\sigma\mu\nu}p_{\mu}\beta_{\nu}\nabla_{\rho}\beta_{\sigma}\biggr{)}\\ &\quad+\biggl{(}-\frac{1}{8}\delta(p^{2})R^{\alpha\rho}\beta_{\alpha}+\frac{1}{4}\delta^{\prime}(p^{2})R^{\rho\sigma\mu\nu}p_{\mu}\beta_{\nu}p_{\sigma}\biggr{)}D_{\rho}f_{(0)}^{\prime}+\biggl{(}\frac{1}{8}\delta(p^{2})R^{\rho\sigma\mu\nu}p_{\mu}\beta_{\nu}\beta_{\sigma}\biggr{)}D_{\rho}f_{(0)}^{\prime\prime}\,,\end{split} (137)
Dμf(0)(12p2Rμαpα112p2Rpμ+23(p2)2Rαβpμpαpβ)=f(0)(13p2pR+23(p2)2pRαβpαpβ)+(12p2Rμαpα)Dμf(0),\begin{split}&D_{\mu}f_{(0)}\biggl{(}-\frac{1}{2p^{2}}R^{\mu\alpha}p_{\alpha}-\frac{1}{12p^{2}}R\,p^{\mu}+\frac{2}{3(p^{2})^{2}}R^{\alpha\beta}p^{\mu}p_{\alpha}p_{\beta}\biggr{)}\\ &=f_{(0)}\biggl{(}-\frac{1}{3p^{2}}p\cdot\nabla R+\frac{2}{3(p^{2})^{2}}p\cdot\nabla R^{\alpha\beta}p_{\alpha}p_{\beta}\biggr{)}+\biggl{(}-\frac{1}{2p^{2}}R^{\mu\alpha}p_{\alpha}\biggr{)}D_{\mu}f_{(0)}\,,\end{split} (138)
Dμf(0)(124Rμαβα+112p2Rαβγμpαββpγ)=f(0)(148βR112p2pRαβpαββ+112p2βRαβpαpβ)+(124Rμαβα+112p2Rαβγμpαββpγ)Dμf(0),\begin{split}&D_{\mu}f_{(0)}^{\prime}\biggl{(}-\frac{1}{24}R^{\mu\alpha}\beta_{\alpha}+\frac{1}{12p^{2}}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}p_{\gamma}\biggr{)}\\ &=f_{(0)}^{\prime}\biggl{(}-\frac{1}{48}\beta\cdot\nabla R-\frac{1}{12p^{2}}p\cdot\nabla R^{\alpha\beta}p_{\alpha}\beta_{\beta}+\frac{1}{12p^{2}}\beta\cdot\nabla R^{\alpha\beta}p_{\alpha}p_{\beta}\biggr{)}\\ &\quad+\biggl{(}-\frac{1}{24}R^{\mu\alpha}\beta_{\alpha}+\frac{1}{12p^{2}}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}p_{\gamma}\biggr{)}D_{\mu}f_{(0)}^{\prime}\,,\end{split} (139)
Dμf(0)′′(124Rαβγμpαβββγ112p2Rαβγδpμpαpγβββδ)=f(0)′′(124pRαββαββ124βRαβpαββ112p2pRαβγδpαpγβββδ+116Rαβμνpαββμβν)+(16p2Rαβγμpαpγββ)Dμf(0)+(124Rαβγμpαβββγ)Dμf(0)′′,\begin{split}&D_{\mu}f_{(0)}^{\prime\prime}\biggl{(}-\frac{1}{24}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}\beta_{\gamma}-\frac{1}{12p^{2}}R_{\alpha\beta\gamma\delta}p^{\mu}p^{\alpha}p^{\gamma}\beta^{\beta}\beta^{\delta}\biggr{)}\\ &=f_{(0)}^{\prime\prime}\biggl{(}\frac{1}{24}p\cdot\nabla R^{\alpha\beta}\beta_{\alpha}\beta_{\beta}-\frac{1}{24}\beta\cdot\nabla R^{\alpha\beta}p_{\alpha}\beta_{\beta}-\frac{1}{12p^{2}}p\cdot\nabla R_{\alpha\beta\gamma\delta}p^{\alpha}p^{\gamma}\beta^{\beta}\beta^{\delta}\\ &\qquad+\frac{1}{16}R^{\alpha\beta\mu\nu}p_{\alpha}\beta_{\beta}\nabla_{\mu}\beta_{\nu}\biggr{)}+\biggl{(}\frac{1}{6p^{2}}R^{\alpha\beta\gamma\mu}p_{\alpha}p_{\gamma}\beta_{\beta}\biggr{)}D_{\mu}f_{(0)}^{\prime}+\biggl{(}-\frac{1}{24}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}\beta_{\gamma}\biggr{)}D_{\mu}f_{(0)}^{\prime\prime}\,,\end{split} (140)
14Dμ[ρβμ]pν[ρβν]f(0)′′=f(0)′′(18Rαβμνpαββ[μβν])+(14Rαββα)Dβf(0)+(14[ρβμ]pν[ρβν])Dμf(0)′′,\begin{split}&-\frac{1}{4}D_{\mu}\nabla^{[\rho}\beta^{\mu]}p^{\nu}\nabla_{[\rho}\beta_{\nu]}f_{(0)}^{\prime\prime}\\ &=f_{(0)}^{\prime\prime}\biggl{(}-\frac{1}{8}R_{\alpha\beta\mu\nu}p^{\alpha}\beta^{\beta}\nabla^{[\mu}\beta^{\nu]}\biggr{)}+\biggl{(}\frac{1}{4}R^{\alpha\beta}\beta_{\alpha}\biggr{)}D_{\beta}f_{(0)}^{\prime}+\biggl{(}-\frac{1}{4}\nabla^{[\rho}\beta^{\mu]}p^{\nu}\nabla_{[\rho}\beta_{\nu]}\biggr{)}D_{\mu}f_{(0)}^{\prime\prime}\,,\end{split} (141)

and

Dμpμ4p2pν[ρβν]pσ[ρβσ]f(0)′′=(12p2Rαβγμpαββpγ)Dμf(0).\begin{split}D_{\mu}\frac{p^{\mu}}{4p^{2}}p_{\nu}\nabla^{[\rho}\beta^{\nu]}p^{\sigma}\nabla_{[\rho}\beta_{\sigma]}f_{(0)}^{\prime\prime}&=\biggl{(}-\frac{1}{2p^{2}}R^{\alpha\beta\gamma\mu}p_{\alpha}\beta_{\beta}p_{\gamma}\biggr{)}D_{\mu}f_{(0)}^{\prime}\,.\end{split} (142)

Collecting them, we obtain Eq. (67).

Appendix C Integration formulas

Here, we present several Integration formulas. We first define

Cn:=12π20𝑑ρρn[nF(ρμ)(1)nnF(ρ+μ)],nF(z):=1eβz+1.C_{n}:=\frac{1}{2\pi^{2}}\int_{0}^{\infty}d\rho\rho^{n}\Bigl{[}n_{F}(\rho-\mu)-(-1)^{n}n_{F}(\rho+\mu)\Bigr{]}\,,\quad n_{F}(z):=\frac{1}{e^{\beta z}+1}\,. (143)

In particular, the first four CnC_{n}’s are

C0=μ2π2,\displaystyle\displaystyle C_{0}=\frac{\mu}{2\pi^{2}}\,, (144)
C1=μ24π2+T212,\displaystyle\displaystyle C_{1}=\frac{\mu^{2}}{4\pi^{2}}+\frac{T^{2}}{12}\,, (145)
C2=20μ𝑑νC1(ν)=μ36π2+μT26,\displaystyle\displaystyle C_{2}=2\int_{0}^{\mu}d\nu\,C_{1}(\nu)=\frac{\mu^{3}}{6\pi^{2}}+\frac{\mu T^{2}}{6}\,, (146)
C3=30μ𝑑νC2(ν)=μ48π2+μ2T24.\displaystyle\displaystyle C_{3}=3\int_{0}^{\mu}d\nu\,C_{2}(\nu)=\frac{\mu^{4}}{8\pi^{2}}+\frac{\mu^{2}T^{2}}{4}\,. (147)

Also in the integral of angular degrees of freedom, we can replace the product of pμp_{\mu}’s in the integral, as follows:

pαp0ξα,pαpβ(p0)2ξαξβ+𝒑23Δαβ,pαpβpγ(p0)3ξαξβξγ+p0𝒑23(ξαΔβγ+ξβΔγα+ξγΔαβ),pαpβpγpδ(p0)4ξαξβξγξδ+(p0)2𝒑23(ξαξβΔγδ+ξαξγΔβδ+ξαξδΔβγ+ξβξγΔαδ+ξβξδΔαγ+ξγξδΔαβ)+|𝒑|415(ΔαβΔγδ+ΔαγΔβδ+ΔαδΔβγ),pαpβpγpδpλ(p0)5ξαξβξγξδξλ+(p0)3𝒑23(Δαβξγξδξλ+Δαγξβξδξλ+Δαδξβξγξλ+Δαλξβξβξγ+Δβγξαξδξλ+Δβδξαξγξλ+Δβλξαξγξδ+Δγδξαξβξλ+Δγλξαξβξδ+Δδλξαξβξγ)+115p0|𝒑|4[ξα(ΔβγΔδλ+ΔβδΔγλ+ΔβλΔγδ)+ξβ(ΔαγΔδλ+ΔαδΔγλ+ΔαλΔγδ)+ξγ(ΔαβΔδλ+ΔαδΔβλ+ΔαλΔβδ)+ξδ(ΔαβΔγλ+ΔαγΔβλ+ΔαλΔβγ)+ξλ(ΔαβΔγδ+ΔαγΔβδ+ΔαδΔβγ)],\begin{split}p_{\alpha}&\to p_{0}\xi_{\alpha},\\ p_{\alpha}p_{\beta}&\to(p_{0})^{2}\xi_{\alpha}\xi_{\beta}+\frac{{\boldsymbol{p}}^{2}}{3}\Delta_{\alpha\beta},\\ p_{\alpha}p_{\beta}p_{\gamma}&\to(p_{0})^{3}\xi_{\alpha}\xi_{\beta}\xi_{\gamma}+\frac{p_{0}{\boldsymbol{p}}^{2}}{3}(\xi_{\alpha}\Delta_{\beta\gamma}+\xi_{\beta}\Delta_{\gamma\alpha}+\xi_{\gamma}\Delta_{\alpha\beta}),\\ p_{\alpha}p_{\beta}p_{\gamma}p_{\delta}&\to(p_{0})^{4}\xi_{\alpha}\xi_{\beta}\xi_{\gamma}\xi_{\delta}\\ &\quad+\frac{(p_{0})^{2}{\boldsymbol{p}}^{2}}{3}(\xi_{\alpha}\xi_{\beta}\Delta_{\gamma\delta}+\xi_{\alpha}\xi_{\gamma}\Delta_{\beta\delta}+\xi_{\alpha}\xi_{\delta}\Delta_{\beta\gamma}+\xi_{\beta}\xi_{\gamma}\Delta_{\alpha\delta}+\xi_{\beta}\xi_{\delta}\Delta_{\alpha\gamma}+\xi_{\gamma}\xi_{\delta}\Delta_{\alpha\beta})\\ &\quad+\frac{|{\boldsymbol{p}}|^{4}}{15}(\Delta_{\alpha\beta}\Delta_{\gamma\delta}+\Delta_{\alpha\gamma}\Delta_{\beta\delta}+\Delta_{\alpha\delta}\Delta_{\beta\gamma}),\\ p_{\alpha}p_{\beta}p_{\gamma}p_{\delta}p_{\lambda}&\to(p_{0})^{5}\xi_{\alpha}\xi_{\beta}\xi_{\gamma}\xi_{\delta}\xi_{\lambda}\\ &\quad+\frac{(p_{0})^{3}{\boldsymbol{p}}^{2}}{3}(\Delta_{\alpha\beta}\xi_{\gamma}\xi_{\delta}\xi_{\lambda}+\Delta_{\alpha\gamma}\xi_{\beta}\xi_{\delta}\xi_{\lambda}+\Delta_{\alpha\delta}\xi_{\beta}\xi_{\gamma}\xi_{\lambda}+\Delta_{\alpha\lambda}\xi_{\beta}\xi_{\beta}\xi_{\gamma}+\Delta_{\beta\gamma}\xi_{\alpha}\xi_{\delta}\xi_{\lambda}\\ &\quad\quad+\Delta_{\beta\delta}\xi_{\alpha}\xi_{\gamma}\xi_{\lambda}+\Delta_{\beta\lambda}\xi_{\alpha}\xi_{\gamma}\xi_{\delta}+\Delta_{\gamma\delta}\xi_{\alpha}\xi_{\beta}\xi_{\lambda}+\Delta_{\gamma\lambda}\xi_{\alpha}\xi_{\beta}\xi_{\delta}+\Delta_{\delta\lambda}\xi_{\alpha}\xi_{\beta}\xi_{\gamma})\\ &\quad+\frac{1}{15}p_{0}|{\boldsymbol{p}}|^{4}\Bigl{[}\xi_{\alpha}(\Delta_{\beta\gamma}\Delta_{\delta\lambda}+\Delta_{\beta\delta}\Delta_{\gamma\lambda}+\Delta_{\beta\lambda}\Delta_{\gamma\delta})\\ &\quad\quad+\xi_{\beta}(\Delta_{\alpha\gamma}\Delta_{\delta\lambda}+\Delta_{\alpha\delta}\Delta_{\gamma\lambda}+\Delta_{\alpha\lambda}\Delta_{\gamma\delta})+\xi_{\gamma}(\Delta_{\alpha\beta}\Delta_{\delta\lambda}+\Delta_{\alpha\delta}\Delta_{\beta\lambda}+\Delta_{\alpha\lambda}\Delta_{\beta\delta})\\ &\quad\quad+\xi_{\delta}(\Delta_{\alpha\beta}\Delta_{\gamma\lambda}+\Delta_{\alpha\gamma}\Delta_{\beta\lambda}+\Delta_{\alpha\lambda}\Delta_{\beta\gamma})+\xi_{\lambda}(\Delta_{\alpha\beta}\Delta_{\gamma\delta}+\Delta_{\alpha\gamma}\Delta_{\beta\delta}+\Delta_{\alpha\delta}\Delta_{\beta\gamma})\Bigr{]}\,,\end{split} (148)

where ξμ:=(1,𝟎)\xi^{\mu}:=(1,{\boldsymbol{0}}) and Δμν:=ξμξνημν\Delta^{\mu\nu}:=\xi^{\mu}\xi^{\nu}-\eta^{\mu\nu}.

Appendix D Alternative derivation of J(2)eqμJ^{\mu}_{{(2)}\text{eq}}

In this appendix, we derive the curvature-induced charge current Jeq(2)μJ_{\text{eq}{(2)}}^{\mu} in Eq. (78), from the thermodynamics of Weyl fermions in a curved spacetime. At the same time, such alternative derivations leading to the same Jeq(2)μJ_{\text{eq}{(2)}}^{\mu} ensures the correctness of the Wigner function (2)μ\mathcal{R}^{\mu}_{(2)} in Eq. (46) and the equilibrium distribution function f(2)f_{(2)} given by Eqs. (65) and (71).

D.1 Diagrammatic computation

First, we derive the current of a chiral fluid under a gravitational field, based on the linear response theory. We consider a Weyl fermion, and the corresponding action is given by

S=i2d4xeη(σaeaμ(μ+iAμ)(μiAμ)σaeaμ)η,\begin{split}S=\frac{i}{2}\int d^{4}xe\eta^{{\dagger}}\Bigl{(}\sigma^{{a}}e_{{~{}a}}^{\mu}(\nabla_{\mu}+iA_{\mu})-(\overleftarrow{\nabla}_{\mu}-iA_{\mu})\sigma^{a}e_{{~{}a}}^{\mu}\Bigr{)}\eta,\end{split} (149)

where we introduce σa=(1,σi)\sigma^{a}=(1,\sigma^{i}) with the Pauli matrices σi(i=1,2,3)\sigma^{i}~{}(i=1,2,3). Here eμa(eaμ)e_{\mu}^{~{}a}(e^{\mu}_{~{}a}) denotes (inverse) vierbein satisfying gμν=eμaeνbηab,ηab=eμaeνbgμνg_{\mu\nu}=e_{\mu}^{~{}a}e_{\nu}^{~{}b}\eta_{ab},~{}\eta^{ab}=e_{\mu}^{~{}a}e_{\nu}^{~{}b}g^{\mu\nu} with the spacetime curved metric gμνg_{\mu\nu} and Minkowski metric ηab=diag(1,1,1,1)\eta_{ab}=\mathrm{diag}(1,-1,-1,-1), and e:=deteμae:=\det e_{\mu}^{{~{}a}}. The left and right covariant derivatives are defined as

μη:=μηi𝒜μη,ημ:=μη+iη𝒜μ,𝒜μ:=12ωμabΣab,Σab:=i4(σ¯aσbσ¯bσa)\begin{split}\nabla_{\mu}\eta&:=\partial_{\mu}\eta-i\mathcal{A}_{\mu}\eta\,,\quad\eta^{{\dagger}}\overleftarrow{\nabla}_{\mu}:=\partial_{\mu}\eta^{\dagger}+i\eta^{\dagger}\mathcal{A}^{{\dagger}}_{\mu}\,,\\ \mathcal{A}_{\mu}&:=\frac{1}{2}\omega_{\mu}^{~{}ab}\Sigma_{ab}\,,\quad\Sigma^{ab}:=\frac{i}{4}(\bar{\sigma}^{a}{\sigma}^{b}-\bar{\sigma}^{b}{\sigma}^{a})\end{split} (150)

with σ¯a:=(1,σi)\bar{\sigma}^{a}:=(1,-\sigma^{i}), which satisfies σ¯aσb+σ¯bσa=σaσ¯b+σbσ¯a=2ηab\bar{\sigma}^{a}{\sigma}^{b}+\bar{\sigma}^{b}{\sigma}^{a}=\sigma^{a}\bar{\sigma}^{b}+\sigma^{b}\bar{\sigma}^{a}=2\eta^{ab}. Furthermore, employing the torsionless condition, we can express the spin connection ωμab=ωμba\omega_{\mu}^{~{}ab}=-\omega_{\mu}^{~{}ba} as

ωμab:=12eνaeρb(CνρμCρνμCμνρ),Cμνρ:=eμc(νeρcρeνc).\begin{split}&\omega_{\mu}^{~{}ab}:=\frac{1}{2}e^{\nu a}e^{\rho b}(C_{\nu\rho\mu}-C_{\rho\nu\mu}-C_{\mu\nu\rho})\,,\\ &C_{\mu\nu\rho}:=e_{\mu}^{~{}c}(\partial_{\nu}e_{\rho c}-\partial_{\rho}e_{\nu c}).\end{split} (151)

The energy-momentum tensor TμνT^{\mu\nu} and U(1)U(1) covariant charge current JμJ^{\mu} are defined as

Tμν=1eδSδeμaeaν=i2η(σμννσμ)η+14ρ(η(σμΣνρ+Σνρσμ)η)gμν,Jμ=1eδSδAμ=ησμη.\begin{split}T^{\mu\nu}&=-\frac{1}{e}\frac{\delta S}{\delta e_{\mu}^{~{}a}}e_{~{}a}^{\nu}=\frac{i}{2}\eta^{\dagger}(\sigma^{\mu}\overrightarrow{\nabla^{\nu}}-\overleftarrow{\nabla^{\nu}}\sigma^{\mu})\eta+\frac{1}{4}\nabla_{\rho}(\eta^{{\dagger}}(\sigma^{\mu}\Sigma^{\nu\rho}+\Sigma^{\nu\rho{\dagger}}\sigma^{\mu})\eta)-\mathcal{L}g^{\mu\nu},\\ J^{\mu}&=-\frac{1}{e}\frac{\delta S}{\delta A_{\mu}}=\eta^{\dagger}\sigma^{\mu}\eta.\end{split} (152)

Note that TμνT^{\mu\nu} is not symmetric, so we introduce the symmetric energy-momentum tensor defined as TSμν:=(Tμν+Tνμ)/2T^{\mu\nu}_{S}:=(T^{\mu\nu}+T^{\nu\mu})/2. In the following, we consider fluctuation around the flat metric gμν=ημν+hμνg_{\mu\nu}=\eta_{\mu\nu}+h_{\mu\nu}.

In the linear response theory, the current in momentum space can be expressed as

Jμ(k)=12Gμνρ(k)hνρ(k)\begin{split}\langle J^{\mu}(k)\rangle=-\frac{1}{2}G^{\mu\nu\rho}(k)h_{\nu\rho}(k)\end{split} (153)

with

Gμνρ(k):=Tnd3k(2π)3ei𝒌𝒙TτJμ(x)TSνρ(0),\begin{split}G^{\mu\nu\rho}(k):=T\sum_{n}\int\frac{d^{3}k}{(2\pi)^{3}}e^{-i{\boldsymbol{k}}\cdot{\boldsymbol{x}}}\langle T_{\tau}J^{\mu}(x)T_{S}^{\nu\rho}(0)\rangle\,,\end{split} (154)

where we define kμ=(0,𝒌)k^{\mu}=(0,{\boldsymbol{k}}) and TτT_{\tau} denotes the imaginary time ordering. The two point correlator is computed with the Feynman rule in flat spacetime:

[Uncaptioned image]=σ¯μpμp2,\displaystyle\displaystyle\parbox{34.14322pt}{\includegraphics[width=34.69038pt]{prop} }=\frac{-\bar{\sigma}^{\mu}p_{\mu}}{p^{2}}\,, (155)
[Uncaptioned image]=σμ,\displaystyle\displaystyle\parbox{48.36958pt}{\includegraphics[width=47.69846pt]{vertex2} }=\sigma^{\mu}\,, (156)
[Uncaptioned image]=14σλ[δλμ(pν+pν)+δλν(pμ+pμ)2ημν(pλ+pλ)].\displaystyle\displaystyle\parbox{51.21504pt}{\includegraphics[width=52.03227pt]{vertex} }=\frac{1}{4}\sigma^{\lambda}\Bigl{[}\delta_{\lambda}^{\mu}(p^{\nu}+p^{\prime\nu})+\delta_{\lambda}^{\nu}(p^{\mu}+p^{\prime\mu})-2\eta^{\mu\nu}(p_{\lambda}+p^{\prime}_{\lambda})\Bigr{]}. (157)

In momentum space, at one-loop order, we get

Gμνρ(k)hνρ=(1)Tn𝒑pαpβp2p2tr[σ¯ασμσ¯βσλ]×hνρ4(δλν(pρ+pρ)+δλρ(pν+pν)2ηνρ(pλ+pλ))=αβγ(k)[ημβ(hγαηαγhρρ)ηβα(hμγημγhρρ)+ημα(hβγηβγhρρ)+iεμβλα(hλγδλγhρρ)],\begin{split}G^{\mu\nu\rho}(k)h_{\nu\rho}&=(-1)T\sum_{n}\int_{\boldsymbol{p}}\frac{p_{\alpha}p^{\prime}_{\beta}}{p^{2}p^{\prime 2}}\,\text{tr}\Bigl{[}\bar{\sigma}^{\alpha}\sigma^{\mu}\bar{\sigma}^{\beta}\sigma^{\lambda}\Bigr{]}\\ &\qquad\qquad\times\frac{h_{\nu\rho}}{4}\Bigl{(}\delta^{\nu}_{\lambda}(p^{\rho}+p^{\prime\rho})+\delta^{\rho}_{\lambda}(p^{\nu}+p^{\prime\nu})-2\eta^{\nu\rho}(p_{\lambda}+p^{\prime}_{\lambda})\Bigr{)}\\ &=-\mathcal{I}_{\alpha\beta\gamma}(k)\Bigl{[}\eta^{\mu\beta}(h^{\gamma\alpha}-\eta^{\alpha\gamma}h_{\rho}^{\rho})-\eta^{\beta\alpha}(h^{\mu\gamma}-\eta^{\mu\gamma}h_{\rho}^{\rho})\\ &\qquad\qquad+\eta^{\mu\alpha}(h^{\beta\gamma}-\eta^{\beta\gamma}h_{\rho}^{\rho})+i\varepsilon^{\mu\beta\lambda\alpha}(h_{\lambda}^{\gamma}-\delta_{\lambda}^{\gamma}h_{\rho}^{\rho})\Bigr{]},\end{split} (158)

where we denote 𝒑=d3p(2π)3\int_{\boldsymbol{p}}=\int\frac{d^{3}p}{(2\pi)^{3}}, p=p+kp^{\prime}=p+k and pμ=(iπT(2n+1)+μ,𝒑)p^{\mu}=(i\pi T(2n+1)+\mu,{\boldsymbol{p}}), and the antisymmetric tensor εμνρσ\varepsilon^{\mu\nu\rho\sigma} is normalized as ε0123=+1\varepsilon^{0123}=+1. Also we introduced

αβγ(k):=Tn𝒑pαpβ(pγ+pγ)p2p2.\mathcal{I}_{\alpha\beta\gamma}(k):=T\sum_{n}\int_{\boldsymbol{p}}\frac{p_{\alpha}p^{\prime}_{\beta}(p_{\gamma}+p^{\prime}_{\gamma})}{p^{2}p^{\prime 2}}\,. (159)

In order to compute the liner response to the gravitational field, we expand αβγ(k)\mathcal{I}_{\alpha\beta\gamma}(k) in terms of kk and define αβγ(n)(k)\mathcal{I}^{(n)}_{\alpha\beta\gamma}(k) to be the O(kn)O(k^{n}) contribution of αβγ(k)\mathcal{I}_{\alpha\beta\gamma}(k). In particular, we find

αβγ(1)(k)=Tn𝒑1(p2)2(pαpβkγ+2pαkβpγ2pαpβpγ2pkp2),αβγ(2)(k)=Tn𝒑(pαpβkγ2pkp22pαkβpγ2pkp2+pαkβkγ2pαpβpγk2p2+8pβpγpα(pk)2(p2)2).\begin{split}\mathcal{I}^{(1)}_{\alpha\beta\gamma}(k)&=T\sum_{n}\int_{\boldsymbol{p}}\frac{1}{(p^{2})^{2}}\Bigl{(}p_{\alpha}p_{\beta}k_{\gamma}+2p_{\alpha}k_{\beta}p_{\gamma}-2p_{\alpha}p_{\beta}p_{\gamma}\frac{2p\cdot k}{p^{2}}\Bigr{)}\,,\\ \mathcal{I}^{(2)}_{\alpha\beta\gamma}(k)&=T\sum_{n}\int_{\boldsymbol{p}}\Bigl{(}-p_{\alpha}p_{\beta}k_{\gamma}\frac{2p\cdot k}{p^{2}}-2p_{\alpha}k_{\beta}p_{\gamma}\frac{2p\cdot k}{p^{2}}\\ &\qquad\qquad\qquad\qquad+p_{\alpha}k_{\beta}k_{\gamma}-2p_{\alpha}p_{\beta}p_{\gamma}\frac{k^{2}}{p^{2}}+8p_{\beta}p_{\gamma}p_{\alpha}\frac{(p\cdot k)^{2}}{(p^{2})^{2}}\Bigr{)}.\end{split} (160)

There are two steps to compute the momentum integrals. First, the radial integral is systematically evaluated with the following formulas:

Fn,m\displaystyle\displaystyle F_{n,m} :=\displaystyle:= Tl𝒑|𝒑|2n2mp02m+1(p2)n+2=F0,02Γ(m+1/2)Γ(mn1/2)Γ(n+2),\displaystyle T\sum_{l}\int_{\boldsymbol{p}}\frac{|{\boldsymbol{p}}|^{2n-2m}p_{0}^{2m+1}}{(p^{2})^{n+2}}=-F_{0,0}\frac{2\Gamma(m+1/2)}{\Gamma(m-n-1/2)\Gamma(n+2)}\,, (161)
F~n,m\displaystyle\displaystyle\tilde{F}_{n,m} :=\displaystyle:= Tl𝒑|𝒑|2n2m(p0)2m(p2)n+1=F~0,0Γ(m1/2)Γ(mn1/2)Γ(n+1)\displaystyle T\sum_{l}\int_{\boldsymbol{p}}\frac{|{\boldsymbol{p}}|^{2n-2m}(p_{0})^{2m}}{(p^{2})^{n+1}}=\tilde{F}_{0,0}\frac{\Gamma(m-1/2)}{\Gamma(m-n-1/2)\Gamma(n+1)} (162)

with

F0,0=18π2μ,\displaystyle\displaystyle F_{0,0}=-\frac{1}{8\pi^{2}}\mu, (163)
F~0,0=18π2(μ2+π23T2).\displaystyle\displaystyle\tilde{F}_{0,0}=\frac{1}{8\pi^{2}}\Bigl{(}\mu^{2}+\frac{\pi^{2}}{3}T^{2}\Bigr{)}\,. (164)

The above formulas are proved in Appendix E. Second, for the angle integrals in Eq. (160), we replace the momentum products pμ1pμjp_{\mu_{1}}\cdots p_{\mu_{j}}, as shown in Eq. (148). Then we obtain

αβγ(1)=(F~1,1+4F~2,13)ξαξβkγ+(2F~1,1+4F~2,13)ξαξγkβ+4F~2,13ξβξγkα+(F~1,03+4F~2,015)Δαβkγ+(2F~1,03+4F~2,015)Δαγkβ+4F~2,015kαΔγβ=F~0,02(ξαξγkβ+ξβξγkαΔαγkβ+Δγβkα),αβγ(2)=(2F1,0+F0,0+16F2,015)ξαkβkγ+(23F1,0+16F2,015)ξβkαkγ+(43F1,0+16F2,015)ξγkαkβ(2F1,1+8F2,13)k2ξαξβξγ(23F1,0+8F2,015)k2(ξαΔβγ+ξβΔγα+ξγΔαβ)=F0,06(ξαkβkγ+ξβkαkγ2ξγkαkβk2ξαξβξγ+k2ξαΔβγ+k2ξβΔγα+k2ξγΔαβ),\begin{split}\mathcal{I}^{(1)}_{\alpha\beta\gamma}&=\Bigl{(}\tilde{F}_{1,1}+4\frac{\tilde{F}_{2,1}}{3}\Bigr{)}\xi_{\alpha}\xi_{\beta}k_{\gamma}+\Bigl{(}2\tilde{F}_{1,1}+4\frac{\tilde{F}_{2,1}}{3}\Bigr{)}\xi_{\alpha}\xi_{\gamma}k_{\beta}+4\frac{\tilde{F}_{2,1}}{3}\xi_{\beta}\xi_{\gamma}k_{\alpha}\\ &\qquad+\Bigl{(}\frac{\tilde{F}_{1,0}}{3}+4\frac{\tilde{F}_{2,0}}{15}\Bigr{)}\Delta_{\alpha\beta}k_{\gamma}+\Bigl{(}2\frac{\tilde{F}_{1,0}}{3}+4\frac{\tilde{F}_{2,0}}{15}\Bigr{)}\Delta_{\alpha\gamma}k_{\beta}+4\frac{\tilde{F}_{2,0}}{15}k_{\alpha}\Delta_{\gamma\beta}\\ &=\frac{\tilde{F}_{0,0}}{2}(-\xi_{\alpha}\xi_{\gamma}k_{\beta}+\xi_{\beta}\xi_{\gamma}k_{\alpha}-\Delta_{\alpha\gamma}k_{\beta}+\Delta_{\gamma\beta}k_{\alpha})\,,\\ \mathcal{I}^{(2)}_{\alpha\beta\gamma}&=\Bigl{(}2F_{1,0}+F_{0,0}+\frac{16F_{2,0}}{15}\Bigr{)}\xi_{\alpha}k_{\beta}k_{\gamma}+\Bigl{(}\frac{2}{3}F_{1,0}+\frac{16F_{2,0}}{15}\Bigr{)}\xi_{\beta}k_{\alpha}k_{\gamma}+\Bigl{(}\frac{4}{3}F_{1,0}+\frac{16F_{2,0}}{15}\Bigr{)}\xi_{\gamma}k_{\alpha}k_{\beta}\\ &\qquad-\Bigr{(}2F_{1,1}+\frac{8F_{2,1}}{3}\Bigl{)}k^{2}\xi_{\alpha}\xi_{\beta}\xi_{\gamma}-\Bigl{(}\frac{2}{3}F_{1,0}+\frac{8F_{2,0}}{15}\Bigr{)}k^{2}\bigl{(}\xi_{\alpha}\Delta_{\beta\gamma}+\xi_{\beta}\Delta_{\gamma\alpha}+\xi_{\gamma}\Delta_{\alpha\beta}\bigr{)}\\ &=\frac{F_{0,0}}{6}\bigl{(}\xi_{\alpha}k_{\beta}k_{\gamma}+\xi_{\beta}k_{\alpha}k_{\gamma}-2\xi_{\gamma}k_{\alpha}k_{\beta}-k^{2}\xi_{\alpha}\xi_{\beta}\xi_{\gamma}+k^{2}\xi_{\alpha}\Delta_{\beta\gamma}+k^{2}\xi_{\beta}\Delta_{\gamma\alpha}+k^{2}\xi_{\gamma}\Delta_{\alpha\beta}\bigr{)}\,,\end{split} (165)

where we denote Δμν=ξμξνημν\Delta_{\mu\nu}=\xi_{\mu}\xi_{\nu}-\eta_{\mu\nu}. As a result, the O(k)O(k) contribution in Eq. (158) is written as

G(1)μνρ(k)hνρ=12F~0,0(ξαξγkβ+ξβξγkαΔαγkβ+Δγβkα)×(ημβ(hγαηαγhρρ)ηβα(hμγημγhρρ)+ημα(hβγηβγhρρ)+iεμβλα(hλγδλγhρρ))=2iε0μjkF~0,0hk0kj,\begin{split}&G_{(1)}^{\mu\nu\rho}(k)h_{\nu\rho}\\ &=-\frac{1}{2}\tilde{F}_{0,0}(-\xi_{\alpha}\xi_{\gamma}k_{\beta}+\xi_{\beta}\xi_{\gamma}k_{\alpha}-\Delta_{\alpha\gamma}k_{\beta}+\Delta_{\gamma\beta}k_{\alpha})\\ &\quad\times\bigl{(}\eta^{\mu\beta}(h^{\gamma\alpha}-\eta^{\alpha\gamma}h_{\rho}^{\rho})-\eta^{\beta\alpha}(h^{\mu\gamma}-\eta^{\mu\gamma}h_{\rho}^{\rho})+\eta^{\mu\alpha}(h^{\beta\gamma}-\eta^{\beta\gamma}h_{\rho}^{\rho})+i\varepsilon^{\mu\beta\lambda\alpha}(h_{\lambda}^{\gamma}-\delta_{\lambda}^{\gamma}h_{\rho}^{\rho})\bigr{)}\\ &=-2i\varepsilon^{0\mu jk}\tilde{F}_{0,0}h_{k}^{0}k_{j}\,,\end{split} (166)

which reproduces the CVE:

J(1)μ=18π2(μ2+π23T2)ε0μjkjhk0=14π2(μ2+π23T2)ωμ\langle J^{\mu}_{(1)}\rangle=-\frac{1}{8\pi^{2}}\biggl{(}\mu^{2}+\frac{\pi^{2}}{3}T^{2}\biggr{)}\varepsilon^{0\mu jk}\partial_{j}h_{k}^{0}\\ =\frac{1}{4\pi^{2}}\biggl{(}\mu^{2}+\frac{\pi^{2}}{3}T^{2}\biggr{)}\omega^{\mu} (167)

with ωμ=εμνρσξνρhρλξλ/2\omega^{\mu}=\varepsilon^{\mu\nu\rho\sigma}\xi_{\nu}\partial_{\rho}h_{\rho\lambda}\xi^{\lambda}/2. Similarly, the O(k2)O(k^{2}) parts are computed as

G(2)μνρ(k)hνρ=13F0,0[h0αkαkμ+hμ0k2+ξμ(hγαkαkγ+2h00k2hααk2)].\begin{split}G_{(2)}^{\mu\nu\rho}(k)h_{\nu\rho}&=-\frac{1}{3}F_{0,0}\Bigl{[}-h^{0\alpha}k_{\alpha}k^{\mu}+h^{\mu 0}k^{2}+\xi^{\mu}(h^{\gamma\alpha}k_{\alpha}k_{\gamma}+2h^{00}k^{2}-h^{\alpha}_{\alpha}k^{2})\Bigr{]}\,.\end{split} (168)

For the stationary gravitational field (0hμν=0\partial_{0}h_{\mu\nu}=0), we eventually derive

J(2)μ=μ48π2[αμh0α2hμ0ξμ(αγhγα2hαα+22h00)]μ24π2[R0μ12ξμR+2ξμR00],\begin{split}\langle J_{(2)}^{\mu}\rangle&=-\frac{\mu}{48\pi^{2}}\Bigl{[}\partial_{\alpha}\partial^{\mu}h^{0\alpha}-\partial^{2}h^{\mu 0}-\xi^{\mu}(\partial_{\alpha}\partial_{\gamma}h^{\gamma\alpha}-\partial^{2}h^{\alpha}_{\alpha}+2\partial^{2}h^{00})\Bigr{]}\\ &\simeq\frac{\mu}{24\pi^{2}}\biggl{[}R^{0\mu}-\frac{1}{2}\xi^{\mu}R+2\xi^{\mu}R^{00}\biggr{]}\,,\end{split} (169)

where we employ

Rμν12(νμhρρνρhρμρμhνρ+ρρhνμ),R2hρρμρhρμ.\begin{split}&R_{\mu\nu}\simeq\frac{1}{2}(\partial_{\nu}\partial_{\mu}h_{\rho}^{\rho}-\partial_{\nu}\partial^{\rho}h_{\rho\mu}-\partial_{\rho}\partial_{\mu}h_{\nu}^{\rho}+\partial_{\rho}\partial^{\rho}h_{\nu\mu})\,,\\ &R\simeq\partial^{2}h_{\rho}^{\rho}-\partial^{\mu}\partial^{\rho}h_{\rho\mu}\,.\end{split} (170)

The above current J(2)μ\langle J_{(2)}^{\mu}\rangle is consistent with J(2)eqμJ^{\mu}_{{(2)}\text{eq}} in Eq. (78).

D.2 Riemann normal coordinate expansion

We reproduce the fermionic current in Eq. (78), by employing the Riemann normal coordinate (RNC) expansion parker2009quantum . We first look for the propagator that satisfies

iγμμxS(x,x)=|g(x)|1/2δ(xx),i\gamma^{\mu}\nabla^{x}_{\mu}S(x,x^{\prime})={|-g(x)|}^{-1/2}\delta(x-x^{\prime})\,, (171)

where we denote g=det(gμν)g=\det(g_{\mu\nu}) and Sab(x,x)=iTψa(x)ψ¯b(x)S_{ab}(x,x^{\prime})=-i\langle T\psi_{a}(x)\bar{\psi}_{b}(x^{\prime})\rangle. Here μx\nabla^{x}_{\mu} is the diffeomorphic and local Lorentz covariant derivative with respect to xx, and the spin connection is defined as

μψ=(μi4ωμabσab)ψ,σab=i2[γa,γb],ωμab=eνa(μebν+Γρμνebρ).\nabla_{\mu}\psi=\biggl{(}\partial_{\mu}-\frac{i}{4}\omega_{\mu ab}\sigma^{ab}\biggr{)}\psi\,,\quad\sigma^{ab}=\frac{i}{2}[\gamma^{a},\gamma^{b}]\,,\quad\omega_{\mu ab}=e_{\nu a}(\partial_{\mu}e_{~{}b}^{\nu}+\Gamma^{\nu}_{\rho\mu}e_{~{}b}^{\rho})\,. (172)

Further we introduce the following bispinor (not scalar) propagator:

iγμμxG(x,x)=S(x,x).i\gamma^{\mu}\nabla^{x}_{\mu}G(x,x^{\prime})=S(x,x^{\prime})\,. (173)

From Eqs. (171) and (173), we find

|g(x)|1/2(μμ+14R)G(x,x)=δ(xx).-|-g(x)|^{1/2}\biggl{(}\nabla^{\mu}\nabla_{\mu}+\frac{1}{4}R\biggr{)}G(x,x^{\prime})=\delta(x-x^{\prime})\,. (174)

Let us now introduce the RNC. We define the normal coordinate yy and the origin is at xx^{\prime}, that is, we replace xyx\to y and x0x^{\prime}\to 0. In order to evaluate above Green’s function, we perform the RNC expansion, as follows:

gμν(x)=ημν+13Rμανβyαyβ+,\displaystyle\displaystyle g_{\mu\nu}(x)=\eta_{\mu\nu}+\frac{1}{3}R_{\mu\alpha\nu\beta}y^{\alpha}y^{\beta}+\cdots\,, (175)
|g(x)|=1+13Rαβyαyβ+,\displaystyle\displaystyle|-g(x)|=1+\frac{1}{3}R_{\alpha\beta}y^{\alpha}y^{\beta}+\cdots\,, (176)
Γμνρ(x)=23Rρ(μν)αyα+,\displaystyle\displaystyle\Gamma^{\rho}_{\mu\nu}(x)=\frac{2}{3}{R^{\rho}}_{(\mu\nu)\alpha}y^{\alpha}+\cdots\,, (177)
eμa(x)=eλa(δμλ+16Rλνμρyνyρ)+,\displaystyle\displaystyle e^{a}_{~{}\mu}(x)=e^{a}_{~{}\lambda}\biggl{(}\delta^{\lambda}_{\mu}+\frac{1}{6}{R^{\lambda}}_{\nu\mu\rho}y^{\nu}y^{\rho}\biggr{)}+\cdots\,, (178)
ωμαβ(x)=12Rαβμνyν+,\displaystyle\displaystyle\omega_{\mu\alpha\beta}(x)=\frac{1}{2}R_{\alpha\beta\mu\nu}y^{\nu}+\cdots\,, (179)

where \cdots denotes the O(R2)O(R^{2}) or O(R)O(\partial R) contribution. Note that all of the above curvature tensors are evaluated at y=0y=0. We thus reduce Eq. (174), as follows:

δ(y)=[ημνμyνy14R16Rαβyαyβy2+13Rμανβyαyβyμyν23Rαβyαyβi4Rμναβσαβyμyν]G(x,x)+.\begin{split}\delta(y)&=\biggl{[}-\eta^{\mu\nu}\partial^{y}_{\mu}\partial^{y}_{\nu}-\frac{1}{4}R-\frac{1}{6}R_{\alpha\beta}y^{\alpha}y^{\beta}\partial_{y}^{2}+\frac{1}{3}R_{\mu\alpha\nu\beta}y^{\alpha}y^{\beta}\partial_{y}^{\mu}\partial_{y}^{\nu}\\ &\qquad-\frac{2}{3}R_{\alpha\beta}y^{\alpha}\partial_{y}^{\beta}-\frac{i}{4}R_{\mu\nu\alpha\beta}\sigma^{\alpha\beta}y^{\mu}\partial_{y}^{\nu}\biggr{]}G(x,x^{\prime})+\cdots\,.\end{split} (180)

Now we perform the Fourier transformation:

G(x,x)=peipyG(p)G(x,x^{\prime})=\int_{p}e^{ip\cdot y}G(p) (181)

with p=d4p(2π)4\int_{p}=\int\frac{d^{4}p}{(2\pi)^{4}}. Then G(p)G(p) obeys

1=[ημνpμpν14R16Rαβpαpβp2+13Rμανβpαpβpμpν+23Rαβpαpβi4Rμναβσαβpμpν+]G(p):=(p2+𝒟)G(p),\begin{split}1&=\biggl{[}\eta^{\mu\nu}p_{\mu}p_{\nu}-\frac{1}{4}R-\frac{1}{6}R_{\alpha\beta}\partial^{\alpha}_{p}\partial^{\beta}_{p}p^{2}+\frac{1}{3}R_{\mu\alpha\nu\beta}\partial^{\alpha}_{p}\partial_{p}^{\beta}p^{\mu}p^{\nu}\\ &\qquad+\frac{2}{3}R_{\alpha\beta}\partial_{p}^{\alpha}p^{\beta}-\frac{i}{4}R_{\mu\nu\alpha\beta}\sigma^{\alpha\beta}\partial_{p}^{\mu}p^{\nu}+\cdots\biggr{]}G(p)\\ &:=\Bigl{(}p^{2}+\mathcal{D}\Bigr{)}G(p)\,,\end{split} (182)

where we denote p2=ημνpμpνp^{2}=\eta^{\mu\nu}p_{\mu}p_{\nu} and 𝒟\mathcal{D} is the derivative operators of O(R)O(R). The above equation is solved sequentially, as follows:

G(p)=1p2[1𝒟G(p)]+=1p2[1𝒟1p2]+=1p2112(p2)2R+23(p2)3Rαβpαpβ+.\begin{split}G(p)&=\frac{1}{p^{2}}\Bigl{[}1-\mathcal{D}G(p)\Bigr{]}+\cdots=\frac{1}{p^{2}}\biggl{[}1-\mathcal{D}\frac{1}{p^{2}}\biggr{]}+\cdots\\ &=\frac{1}{p^{2}}-\frac{1}{12(p^{2})^{2}}R+\frac{2}{3(p^{2})^{3}}R_{\alpha\beta}p^{\alpha}p^{\beta}+\cdots\,.\end{split} (183)

Thus we obtain

S(x,x)=iγμ(x)μpeipyG(p)=peipy(γpp2+12(p2)2Rμνγμpν+γp12(p2)2R2γp3(p2)3Rαβpαpβ)+.\begin{split}S(x,x^{\prime})&=i\gamma^{\mu}(x)\nabla_{\mu}\int_{p}e^{ip\cdot y}G(p)\\ &=\int_{p}e^{ip\cdot y}\,\biggl{(}-\frac{\gamma\cdot p}{p^{2}}+\frac{1}{2(p^{2})^{2}}R_{\mu\nu}\gamma^{\mu}p^{\nu}+\frac{\gamma\cdot p}{12(p^{2})^{2}}R-\frac{2\gamma\cdot p}{3(p^{2})^{3}}R_{\alpha\beta}p^{\alpha}p^{\beta}\biggr{)}+\cdots\,.\end{split} (184)

Performing the Wick rotation, we obtain the vector current as

Jμ=tr[S(x,x)γμ]=Tn𝒑[4pμp22pν(p2)2Rνμpμ3(p2)2R+8pμpαpβ3(p2)3Rαβ]\begin{split}J^{\mu}&=-\text{tr}\Bigl{[}S(x,x)\gamma^{\mu}\Bigr{]}\\ &=T\sum_{n}\int_{\boldsymbol{p}}\biggl{[}\frac{4p^{\mu}}{p^{2}}-\frac{2p^{\nu}}{(p^{2})^{2}}{R_{\nu}}^{\mu}-\frac{p^{\mu}}{3(p^{2})^{2}}R+\frac{8p^{\mu}p^{\alpha}p^{\beta}}{3(p^{2})^{3}}R_{\alpha\beta}\biggr{]}\end{split} (185)

with pμ=(iπT(2n+1)+μ,𝒑)p^{\mu}=(i\pi T(2n+1)+\mu,{\boldsymbol{p}}). The above current is evaluated with Eq. (148). The first term in the above integrand gives the ordinary charge density. The other terms are linear in the curvature tensor, and thus the curvature-induced current JcurvμJ^{\mu}_{\text{curv}} is calculated as

Jcurvμ=2Rμ0F0,013ξμRF0,0+83ξμR00F1,1+ξμ83R00F1,0169Rμ0F1,089ξμRF1,0=2μ24π2[R0μ12ξμR+2ξμR00].\begin{split}J^{\mu}_{\text{curv}}&=-2{R^{\mu}}_{0}F_{0,0}-\frac{1}{3}\xi^{\mu}RF_{0,0}+\frac{8}{3}\xi^{\mu}R_{00}F_{1,1}+\xi^{\mu}\frac{8}{3}R_{00}F_{1,0}-\frac{16}{9}{R^{\mu}}_{0}F_{1,0}-\frac{8}{9}\xi^{\mu}RF_{1,0}\\ &=2\cdot\frac{\mu}{24\pi^{2}}\biggl{[}R^{0\mu}-\frac{1}{2}\xi^{\mu}R+2\xi^{\mu}R^{00}\biggr{]}\,.\end{split} (186)

This is again the same as Eq. (78) up to the factor 22, which comes from the right- and left-handed contributions.

Appendix E Evaluation of Fn,mF_{n,m} and F~n,m\tilde{F}_{n,m}

In this appendix, we derive the formulas of the momentum integrals in Euclidean spacetime, which are applied in Appendix D. We first compute the following integral:

Fn,m\displaystyle F_{n,m} :=\displaystyle:= Tl𝒑|𝒑|2n2mp02m+1(p2)n+2.\displaystyle T\sum_{l}\int_{\boldsymbol{p}}\frac{|{\boldsymbol{p}}|^{2n-2m}p_{0}^{2m+1}}{(p^{2})^{n+2}}\,. (187)

This obeys the recursion relation Fn,m=Fn1,m1+Fn,m1F_{n,m}=F_{n-1,m-1}+F_{n,m-1}, and the solutions are given by

Fn,m=j=0mm!j!(mj)!Fnj,0.\begin{split}F_{n,m}&=\sum_{j=0}^{m}\frac{m!}{j!(m-j)!}F_{n-j,0}\,.\end{split} (188)

We calculate Fn,0F_{n,0} as

Fn,0=Tl𝒑|𝒑|2np0(p2)n+2=TldΩd𝒑2(2π)312(𝒑2)n+1/21(n+1)!(𝒑2)np0(p2)2=(1)nΓ(n+3/2)(n+1)!Γ(3/2)TldΩd𝒑2(2π)312(𝒑2)1/2p0(p2)2=(1)nΓ(n+3/2)Γ(3/2)(n+1)!F0,0.\begin{split}F_{n,0}&=T\sum_{l}\int_{\boldsymbol{p}}\frac{|{\boldsymbol{p}}|^{2n}p_{0}}{(p^{2})^{n+2}}\\ &=T\sum_{l}\int\frac{d\Omega d{\boldsymbol{p}}^{2}}{(2\pi)^{3}}\frac{1}{2}({\boldsymbol{p}}^{2})^{n+1/2}\frac{1}{(n+1)!}\Bigl{(}\frac{\partial}{\partial{\boldsymbol{p}}^{2}}\Bigr{)}^{n}\ \frac{p_{0}}{(p^{2})^{2}}\\ &=\frac{(-1)^{n}\Gamma(n+3/2)}{(n+1)!\Gamma(3/2)}T\sum_{l}\int\frac{d\Omega d{\boldsymbol{p}}^{2}}{(2\pi)^{3}}\frac{1}{2}({\boldsymbol{p}}^{2})^{1/2}\frac{p_{0}}{(p^{2})^{2}}\\ &=\frac{(-1)^{n}\Gamma(n+3/2)}{\Gamma(3/2)(n+1)!}F_{0,0}\,.\end{split} (189)

Therefore, we obtain

Fn,m=F0,0j=0mm!j!(mj)!(1)njΓ(nj+3/2)Γ(3/2)(nj+1)!=F0,02Γ(m+1/2)Γ(mn1/2)Γ(n+2).\begin{split}F_{n,m}&=F_{0,0}\sum_{j=0}^{m}\frac{m!}{j!(m-j)!}\frac{(-1)^{n-j}\Gamma(n-j+3/2)}{\Gamma(3/2)(n-j+1)!}\\ &=-F_{0,0}\frac{2\Gamma(m+1/2)}{\Gamma(m-n-1/2)\Gamma(n+2)}\,.\end{split} (190)

One can check that this solution satisfies the recursion relation:

Fn1,m1+Fn,m1Fn,m=j=0m1(m1)!j!(m1j)!Fn1j,0+j=0m1(m1)!j!(m1j)!Fnj,0j=0mm!j!(mj)!Fnj,0=j=1mm!j!(mj)!jmFnj,0+j=0m1m!j!(mj)!mjmFnj,0j=0mm!j!(mj)!Fnj,0=0.\begin{split}&F_{n-1,m-1}+F_{n,m-1}-F_{n,m}\\ &=\sum_{j=0}^{m-1}\frac{(m-1)!}{j!(m-1-j)!}F_{n-1-j,0}+\sum_{j=0}^{m-1}\frac{(m-1)!}{j!(m-1-j)!}F_{n-j,0}-\sum_{j=0}^{m}\frac{m!}{j!(m-j)!}F_{n-j,0}\\ &=\sum_{j=1}^{m}\frac{m!}{j!(m-j)!}\frac{j}{m}F_{n-j,0}+\sum_{j=0}^{m-1}\frac{m!}{j!(m-j)!}\frac{m-j}{m}F_{n-j,0}-\sum_{j=0}^{m}\frac{m!}{j!(m-j)!}F_{n-j,0}\\ &=0.\end{split} (191)

The overall factor F0,0F_{0,0} in Eq. (190) is computed as

F0,0=Tl𝒑12|𝒑||𝒑|p0(p2)=14π2Tl0d|𝒑|p0p2=14π2Tl0d|𝒑|12(1p0|𝒑|+1p0+|𝒑|)=18π20d|𝒑|(1eβ(|𝒑|μ)+11eβ(|𝒑|+μ)+1)=μ8π2.\begin{split}F_{0,0}&=T\sum_{l}\int_{\boldsymbol{p}}\frac{1}{2|{\boldsymbol{p}}|}\frac{\partial}{\partial|{\boldsymbol{p}}|}\frac{p_{0}}{(p^{2})}\\ &=-\frac{1}{4\pi^{2}}T\sum_{l}\int_{0}^{\infty}d|{\boldsymbol{p}}|\frac{p_{0}}{p^{2}}\\ &=-\frac{1}{4\pi^{2}}T\sum_{l}\int_{0}^{\infty}d|{\boldsymbol{p}}|\frac{1}{2}\biggl{(}\frac{1}{p_{0}-|{\boldsymbol{p}}|}+\frac{1}{p_{0}+|{\boldsymbol{p}}|}\biggr{)}\\ &=-\frac{1}{8\pi^{2}}\int_{0}^{\infty}d|{\boldsymbol{p}}|\biggl{(}\frac{1}{e^{\beta(|{\boldsymbol{p}}|-\mu)}+1}-\frac{1}{e^{\beta(|{\boldsymbol{p}}|+\mu)}+1}\biggr{)}\\ &=-\frac{\mu}{8\pi^{2}}.\end{split} (192)

Also we evaluate

F~n,m:=Tl𝒑(𝒑2)nm(p0)2m(p2)1+n,\begin{split}\tilde{F}_{n,m}:=T\sum_{l}\int_{\boldsymbol{p}}\frac{({\boldsymbol{p}}^{2})^{n-m}(p_{0})^{2m}}{(p^{2})^{1+n}},\end{split} (193)

which obeys the same recursion relation F~n,m=F~n1,m1+F~n,m1\tilde{F}_{n,m}=\tilde{F}_{n-1,m-1}+\tilde{F}_{n,m-1}. In the same manner for Fn,mF_{n,m}, we get

F~n,m\displaystyle\tilde{F}_{n,m} =\displaystyle= j=0mm!j!(mj)!F~nj,0,\displaystyle\sum_{j=0}^{m}\frac{m!}{j!(m-j)!}\tilde{F}_{n-j,0}, (194)
F~n,0\displaystyle\tilde{F}_{n,0} =\displaystyle= (1)nΓ(n+3/2)Γ(3/2)n!F~0,0,\displaystyle\frac{(-1)^{n}\Gamma(n+3/2)}{\Gamma(3/2)n!}\tilde{F}_{0,0}, (195)
F~n,m\displaystyle\tilde{F}_{n,m} =\displaystyle= F~0,0j=0m(1)njm!j!(mj)!Γ(nj+3/2)Γ(3/2)(nj)!\displaystyle\tilde{F}_{0,0}\sum_{j=0}^{m}(-1)^{n-j}\frac{m!}{j!(m-j)!}\frac{\Gamma(n-j+3/2)}{\Gamma(3/2)(n-j)!} (196)
=\displaystyle= F~0,0Γ(m1/2)Γ(mn1/2)Γ(n+1).\displaystyle\tilde{F}_{0,0}\frac{\Gamma(m-1/2)}{\Gamma(m-n-1/2)\Gamma(n+1)}.

The overall factor F~0,0\tilde{F}_{0,0} is calculated as

F~0,0=12π2Tl0d|𝒑||𝒑|2p2=12π2Tl0d|𝒑||𝒑|2(1p0|𝒑|1p0+|𝒑|)=14π20d|𝒑||𝒑|(1eβ(|𝒑|μ)+1+1eβ(|𝒑|+μ)+11)=18π2(μ2+π23T2)+(const).\begin{split}\tilde{F}_{0,0}&=\frac{1}{2\pi^{2}}T\sum_{l}\int_{0}^{\infty}d|{\boldsymbol{p}}|\frac{|{\boldsymbol{p}}|^{2}}{p^{2}}\\ &=\frac{1}{2\pi^{2}}T\sum_{l}\int_{0}^{\infty}d|{\boldsymbol{p}}|\frac{|{\boldsymbol{p}}|}{2}\biggl{(}\frac{1}{p_{0}-|{\boldsymbol{p}}|}-\frac{1}{p_{0}+|{\boldsymbol{p}}|}\biggr{)}\\ &=\frac{1}{4\pi^{2}}\int_{0}^{\infty}d|{\boldsymbol{p}}||{\boldsymbol{p}}|\biggl{(}\frac{1}{e^{\beta(|{\boldsymbol{p}}|-\mu)}+1}+\frac{1}{e^{\beta(|{\boldsymbol{p}}|+\mu)}+1}-1\biggr{)}\\ &=\frac{1}{8\pi^{2}}\left(\mu^{2}+\frac{\pi^{2}}{3}T^{2}\right){\rm+(const)}.\end{split} (197)

Here, (const) denotes the divergent term that is independent of TT and μ\mu.

Appendix F Wigner function under the dynamical gravity (94) and (95)

In this appendix, we derive the Wigner function under a time-dependent gravitational field, Eqs. (94) and (95). Plugging μ\mathcal{R}^{\mu} and PμP_{\mu} given by Eqs. (46) and (11), we write down the kinetic equation (8) as

0=δ(p2)[pD+DμΣnμνDν]f+δ(p2)2Dμ(ΣuμνΣnμν)Dνf~(1)+2p2Dμ[pμQp+2pν(T[μpν]+Sαμνpα)]δ(p2)f+2δ(p2)2p2Dμ[εμνρσpνDρΣnσλDλΣμνu(R~αβνρpρpαβp+2pDΣnνρDρ)]f+2(18λRμνpλpν124λRρσμνpλpνpσpρ+18RρσμνpνpσDρ)pμδ(p2)f.\begin{split}0&=\delta(p^{2})\Bigl{[}p\cdot D+\hbar D_{\mu}\Sigma^{\mu\nu}_{n}D_{\nu}\Bigr{]}f+\delta(p^{2})\hbar^{2}D_{\mu}(\Sigma^{\mu\nu}_{u}-\Sigma^{\mu\nu}_{n})D_{\nu}\tilde{f}_{{(1)}}\\ &\quad+\frac{\hbar^{2}}{p^{2}}D^{\mu}\Bigl{[}-p_{\mu}Q\cdot p+2p^{\nu}\bigl{(}T_{[\mu}p_{\nu]}+S_{\alpha\mu\nu}p^{\alpha}\bigr{)}\Bigr{]}\delta(p^{2})f\\ &\quad+\frac{\hbar^{2}\delta(p^{2})}{2p^{2}}D^{\mu}\biggl{[}\varepsilon_{\mu\nu\rho\sigma}p^{\nu}D^{\rho}\Sigma_{n}^{\sigma\lambda}D_{\lambda}-\Sigma_{\mu\nu}^{u}\bigl{(}{\tilde{R}}^{\alpha\beta\nu\rho}p_{\rho}p_{\alpha}\partial_{\beta}^{p}+2p\cdot D\Sigma_{n}^{\nu\rho}D_{\rho}\bigr{)}\biggr{]}f\\ &\quad+\hbar^{2}\biggl{(}-\frac{1}{8}\nabla_{\lambda}R_{\mu\nu}\partial_{p}^{\lambda}\partial_{p}^{\nu}-\frac{1}{24}\nabla_{\lambda}{R^{\rho}}_{\sigma\mu\nu}\partial_{p}^{\lambda}\partial_{p}^{\nu}\partial_{p}^{\sigma}p_{\rho}+\frac{1}{8}{R^{\rho}}_{\sigma\mu\nu}\partial_{p}^{\nu}\partial_{p}^{\sigma}D_{\rho}\biggr{)}p^{\mu}\delta(p^{2})f\,.\end{split} (198)

Here f~(0)\tilde{f}_{(0)} and f~(1)\tilde{f}_{(1)} involved in ff have already been obtained in Eqs. (86) and (87). After some computation keeping O(hμν)O(h_{\mu\nu}) together with 2pDfO(3)\hbar^{2}p\cdot Df\sim O(\hbar^{3}) and DμfO(hμν)D_{\mu}f\sim O(h_{\mu\nu}), we reduce the kinetic equation (198) to

δ(p2)[(1+1p2hμνpμpν)phμνpμν+Γμνρpμpρpν+(12ΣnμνRαβμνpαpβ)+2(124pRαβpαpβ124pμpρpνpσpRρσμνΣμνunλ2pnμR~αβνλpαβp)]f=0,\begin{split}&\delta(p^{2})\biggl{[}\biggl{(}1+\frac{1}{p^{2}}h^{\mu\nu}p_{\mu}p_{\nu}\biggr{)}p\cdot\partial-h^{\mu\nu}p_{\mu}\partial_{\nu}+\Gamma^{\rho}_{\mu\nu}p^{\mu}p_{\rho}\partial_{p}^{\nu}+\hbar\biggl{(}-\frac{1}{2}\Sigma^{\mu\nu}_{n}R_{\alpha\beta\mu\nu}p^{\alpha}\partial^{\beta}_{p}\biggr{)}\\ &\ +\hbar^{2}\biggl{(}-\frac{1}{24}p\cdot\nabla R_{\alpha\beta}\partial_{p}^{\alpha}\partial_{p}^{\beta}-\frac{1}{24}p^{\mu}p^{\rho}\partial_{p}^{\nu}\partial_{p}^{\sigma}\partial_{p}\cdot\nabla R_{\rho\sigma\mu\nu}-\Sigma_{\mu\nu}^{u}\frac{n_{\lambda}}{2p\cdot n}\nabla^{\mu}\tilde{R}^{\alpha\beta\nu\lambda}p_{\alpha}\partial^{p}_{\beta}\biggr{)}\biggr{]}f=0\,,\end{split} (199)

which yields the second order fluctuation as

f~(2)=Σμνunλ2pnkμkpR~αβνλpαββfflat+124Rαββαββfflat′′+kβ24kppμpρβνβσRρσμνfflat′′′.\begin{split}\tilde{f}_{{(2)}}&=\Sigma_{\mu\nu}^{u}\frac{n_{\lambda}}{2p\cdot n}\frac{k^{\mu}}{k\cdot p}\tilde{R}^{\alpha\beta\nu\lambda}p_{\alpha}\beta_{\beta}f_{\text{flat}}^{\prime}+\frac{1}{24}R_{\alpha\beta}\beta^{\alpha}\beta^{\beta}f_{\text{flat}}^{\prime\prime}+\frac{k\cdot\beta}{24k\cdot p}p^{\mu}p^{\rho}\beta^{\nu}\beta^{\sigma}R_{\rho\sigma\mu\nu}f_{\text{flat}}^{\prime\prime\prime}\,.\end{split} (200)

Collecting Eqs. (86) and (87), the above f~(0),(1),(2)\tilde{f}_{{(0)},{(1)},{(2)}} and the general form of the Wigner function (42), we find

(2)μ/(2π)=δ(p2)[pμ(Σηνunλ2pnkηkpR~αβνλpαββfflat+124Rαββαββfflat′′+kβ24kppηpρβνβσRρσηνfflat′′′)12ikpΣuμνΣnλη(ikν)Rαβληpαββfflat+12p2εμνρσpνΣσλnDρDλ(fflat+f~(0))Σuμνnσ2pnR~αβνσpαββfflat]+1p2[pμQp+2pν(T[μpν]+Sαμνpα)]δ(p2)fflat.\begin{split}&\mathcal{R}^{\mu}_{(2)}/(2\pi)\\ &=\delta(p^{2})\biggl{[}p^{\mu}\biggl{(}\Sigma_{\eta\nu}^{u}\frac{n_{\lambda}}{2p\cdot n}\frac{k^{\eta}}{k\cdot p}\tilde{R}^{\alpha\beta\nu\lambda}p_{\alpha}\beta_{\beta}f_{\text{flat}}^{\prime}+\frac{1}{24}R_{\alpha\beta}\beta^{\alpha}\beta^{\beta}f_{\text{flat}}^{\prime\prime}+\frac{k\cdot\beta}{24k\cdot p}p^{\eta}p^{\rho}\beta^{\nu}\beta^{\sigma}R_{\rho\sigma\eta\nu}f_{\text{flat}}^{\prime\prime\prime}\biggr{)}\\ &\quad-\frac{1}{2ik\cdot p}\Sigma^{\mu\nu}_{u}\Sigma^{\lambda\eta}_{n}(-ik_{\nu})R_{\alpha\beta\lambda\eta}p^{\alpha}\beta^{\beta}f_{\text{flat}}^{\prime}+\frac{1}{2p^{2}}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}\Sigma_{\sigma\lambda}^{n}D_{\rho}D^{\lambda}(f_{\text{flat}}+\tilde{f}_{(0)})\\ &\quad-\Sigma^{\mu\nu}_{u}\frac{n^{\sigma}}{2p\cdot n}\tilde{R}_{\alpha\beta\nu\sigma}p^{\alpha}\beta^{\beta}f_{\text{flat}}^{\prime}\biggr{]}+\frac{1}{p^{2}}\Bigl{[}-p^{\mu}Q\cdot p+2p_{\nu}\bigl{(}T^{[\mu}p^{\nu]}+S^{\alpha\mu\nu}p_{\alpha}\bigr{)}\Bigr{]}\delta(p^{2})f_{\text{flat}}\,.\end{split} (201)

In the above equation, there are the four frame-dependent terms. However, the dependence is totally cancelled out, as shown in the following. These are rewritten as

pμΣηνunλ2pnkηkpR~αβνλpαββfflat\displaystyle p^{\mu}\Sigma_{\eta\nu}^{u}\frac{n_{\lambda}}{2p\cdot n}\frac{k^{\eta}}{k\cdot p}\tilde{R}^{\alpha\beta\nu\lambda}p_{\alpha}\beta_{\beta}f_{\text{flat}}^{\prime} =\displaystyle= (12pμΣναunλpnR~αβνλββ+12pμkηkpΣηαuΣνλnRαβνλββ\displaystyle\biggl{(}-\frac{1}{2}p^{\mu}\Sigma_{\nu\alpha}^{u}\frac{n_{\lambda}}{p\cdot n}\tilde{R}^{\alpha\beta\nu\lambda}\beta_{\beta}+\frac{1}{2}p^{\mu}\frac{k^{\eta}}{k\cdot p}\Sigma_{\eta\alpha}^{u}\Sigma_{\nu\lambda}^{n}{R}^{\alpha\beta\nu\lambda}\beta_{\beta} (202)
+14ετμναpτnλpnR~αβνλββ+14pαεητμνpτnλpnkηkpR~αβνλββ\displaystyle+\frac{1}{4}\varepsilon^{\tau\mu\nu\alpha}p_{\tau}\frac{n^{\lambda}}{p\cdot n}\tilde{R}_{\alpha\beta\nu\lambda}\beta^{\beta}+\frac{1}{4}p^{\alpha}\varepsilon^{\eta\tau\mu\nu}p_{\tau}\frac{n^{\lambda}}{p\cdot n}\frac{k_{\eta}}{k\cdot p}\tilde{R}_{\alpha\beta\nu\lambda}\beta^{\beta}
+14εαητμpτkηkpΣnνλRαβνλββ)fflat,\displaystyle+\frac{1}{4}\varepsilon^{\alpha\eta\tau\mu}p_{\tau}\frac{k_{\eta}}{k\cdot p}\Sigma^{\nu\lambda}_{n}R_{\alpha\beta\nu\lambda}\beta^{\beta}\biggr{)}f_{\text{flat}}^{\prime}\,,
12kpΣuμνΣnληkνRαβληpαββfflat\displaystyle\frac{1}{2k\cdot p}\Sigma^{\mu\nu}_{u}\Sigma^{\lambda\eta}_{n}k_{\nu}R_{\alpha\beta\lambda\eta}p^{\alpha}\beta^{\beta}f_{\text{flat}}^{\prime} =\displaystyle= (12kpΣuναpμΣnληkνRαβληββ+12ΣuμαΣnληRαβληββ\displaystyle\biggl{(}-\frac{1}{2k\cdot p}\Sigma_{u}^{\nu\alpha}p^{\mu}\Sigma^{\lambda\eta}_{n}k_{\nu}R_{\alpha\beta\lambda\eta}\beta^{\beta}+\frac{1}{2}\Sigma^{\mu\alpha}_{u}\Sigma^{\lambda\eta}_{n}R_{\alpha\beta\lambda\eta}\beta^{\beta} (203)
14kpεναμρpρΣnληkνRαβληββ)fflat,\displaystyle-\frac{1}{4k\cdot p}\varepsilon^{\nu\alpha\mu\rho}p_{\rho}\Sigma^{\lambda\eta}_{n}k_{\nu}R_{\alpha\beta\lambda\eta}\beta^{\beta}\biggr{)}f_{\text{flat}}^{\prime},\qquad
Σuμνnσ2pnR~αβνσpαββfflat\displaystyle-\Sigma^{\mu\nu}_{u}\frac{n^{\sigma}}{2p\cdot n}\tilde{R}_{\alpha\beta\nu\sigma}p^{\alpha}\beta^{\beta}f_{\text{flat}}^{\prime} =\displaystyle= (pμΣuναnσ2pnR~αβνσββ+14εναμρpρnσpnR~αβνσββ\displaystyle\biggl{(}p^{\mu}\Sigma_{u}^{\nu\alpha}\frac{n^{\sigma}}{2p\cdot n}\tilde{R}_{\alpha\beta\nu\sigma}\beta^{\beta}+\frac{1}{4}\varepsilon^{\nu\alpha\mu\rho}p_{\rho}\frac{n^{\sigma}}{p\cdot n}\tilde{R}_{\alpha\beta\nu\sigma}\beta^{\beta} (204)
12ΣuμαΣnληRαβληββ)fflat,\displaystyle-\frac{1}{2}\Sigma^{\mu\alpha}_{u}\Sigma^{\lambda\eta}_{n}R_{\alpha\beta\lambda\eta}\beta^{\beta}\biggr{)}f_{\text{flat}}^{\prime}\,,
εμνρσ2p2pνΣσλnDρDλ(fflat+f~(0))\displaystyle\frac{\varepsilon^{\mu\nu\rho\sigma}}{2p^{2}}p_{\nu}\Sigma_{\sigma\lambda}^{n}D_{\rho}D^{\lambda}(f_{\text{flat}}+\tilde{f}_{(0)}) =\displaystyle= (14p2εμνρσpνpη1kpR~αβσηkρpαββ\displaystyle\biggl{(}-\frac{1}{4p^{2}}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}p^{\eta}\frac{1}{k\cdot p}\tilde{R}_{\alpha\beta\sigma\eta}k_{\rho}p^{\alpha}\beta^{\beta} (205)
+14εμνρσpνnηpn1kpR~αβσηkρpαββ)fflat,\displaystyle+\frac{1}{4}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}\frac{n^{\eta}}{p\cdot n}\frac{1}{k\cdot p}\tilde{R}_{\alpha\beta\sigma\eta}k_{\rho}p^{\alpha}\beta^{\beta}\biggr{)}f_{\text{flat}}^{\prime}\,,

where we use Eq. (45), DρDλf(ikρ)(ikλ)f~(0)+(ikρ)ΓλκτpτβκfflatD_{\rho}D_{\lambda}f\simeq(-ik_{\rho})(-ik_{\lambda})\tilde{f}_{(0)}+(-ik_{\rho})\Gamma^{\tau}_{\lambda\kappa}p_{\tau}\beta^{\kappa}f_{\text{flat}}^{\prime} and the second Bianchi identity (133) for Rαβρ[λkτ]R_{\alpha\beta\rho[\lambda}k_{\tau]}. Hence, the four frame-dependent terms in Eq. (LABEL:eq:Rmu2_dyn_app) are recast into

(pμΣηνunλ2pnkηkpR~αβνλpαββ+12kpΣuμνΣnληkνRαβληpαββΣuμνnσ2pnR~αβνσpαββ)fflat+12p2εμνρσpνΣσλnDρDλ(fflat+f~(0))=14p2εμνρσpηpνkρkpR~αβησpαββfflat=14p2Rαβμνpνpαββfflat+14p2Rβνpμpνββfflat14Rβμββfflat14p2kβkpRανpμpνpαfflat+14kβkpRαμpαfflat.\begin{split}&\quad\biggl{(}p^{\mu}\Sigma_{\eta\nu}^{u}\frac{n_{\lambda}}{2p\cdot n}\frac{k^{\eta}}{k\cdot p}\tilde{R}^{\alpha\beta\nu\lambda}p_{\alpha}\beta_{\beta}+\frac{1}{2k\cdot p}\Sigma^{\mu\nu}_{u}\Sigma^{\lambda\eta}_{n}k_{\nu}R_{\alpha\beta\lambda\eta}p^{\alpha}\beta^{\beta}-\Sigma^{\mu\nu}_{u}\frac{n^{\sigma}}{2p\cdot n}\tilde{R}_{\alpha\beta\nu\sigma}p^{\alpha}\beta^{\beta}\biggr{)}f_{\text{flat}}^{\prime}\\ &\quad+\frac{1}{2p^{2}}\varepsilon^{\mu\nu\rho\sigma}p_{\nu}\Sigma_{\sigma\lambda}^{n}D_{\rho}D^{\lambda}(f_{\text{flat}}+\tilde{f}_{(0)})\\ &=\frac{1}{4p^{2}}\varepsilon^{\mu\nu\rho\sigma}p^{\eta}p_{\nu}\frac{k_{\rho}}{k\cdot p}\tilde{R}_{\alpha\beta\eta\sigma}p^{\alpha}\beta^{\beta}f_{\text{flat}}^{\prime}\\ &=-\frac{1}{4p^{2}}{R_{\alpha\beta}}^{\mu\nu}p_{\nu}p^{\alpha}\beta^{\beta}f_{\text{flat}}^{\prime}+\frac{1}{4p^{2}}{R_{\beta}}^{\nu}p^{\mu}p_{\nu}\beta^{\beta}f_{\text{flat}}^{\prime}-\frac{1}{4}{R_{\beta}}^{\mu}\beta^{\beta}f_{\text{flat}}^{\prime}\\ &\quad-\frac{1}{4p^{2}}\frac{k\cdot\beta}{k\cdot p}{R_{\alpha}}^{\nu}p^{\mu}p_{\nu}p^{\alpha}f_{\text{flat}}^{\prime}+\frac{1}{4}\frac{k\cdot\beta}{k\cdot p}{R_{\alpha}}^{\mu}p^{\alpha}f_{\text{flat}}^{\prime}\,.\end{split} (206)

Inserting this into Eq. (LABEL:eq:Rmu2_dyn_app), we finally derive Eqs. (94) and (95).

Appendix G Angle integrals

In this appendix, we derive the integral formulas in terms of the momentum angle valuables. We introduce the following function for the angular integral:

Ij1,jn(x)=xdΩ4πp^j1p^jnx𝒌^𝒑^,\displaystyle\displaystyle I^{j_{1}\cdots,j_{n}}(x)=x\int\frac{d\Omega}{4\pi}\frac{{\hat{p}}^{j_{1}}\cdots{\hat{p}}^{j_{n}}}{x-{\hat{\boldsymbol{k}}}\cdot{\hat{\boldsymbol{p}}}}\,, (207)

where xx involves the positive infinitesimal imaginary part +iη+i\eta and we define x=k0/|𝒌|x=k_{0}/|{\boldsymbol{k}}|. By definition, we can readily show Eqs. (101)-(104) and Eqs. (106)-(108). Let us now evaluate the angle integrals. We define θ\theta and ϕ\phi as the polar and azimuthal angles when the polar axis is along 𝒌^{\hat{\boldsymbol{k}}}. First, we compute

I(x)=x211dyxy=x2lnx+1x1=x2ln|x+1x1|xiπ2θ(1|x|)\begin{split}I(x)=\frac{x}{2}\int_{-1}^{1}\frac{dy}{x-y}&=\frac{x}{2}\ln\frac{x+1}{x-1}=\frac{x}{2}\ln\biggl{|}\frac{x+1}{x-1}\biggr{|}-x\frac{i\pi}{2}\theta(1-|x|)\end{split} (208)

with y=cosθy=\cos\theta. In order to evaluate the other integrals, we prepare the following formulas:

02πdϕ2πp^k=k^ky,02πdϕ2πp^jp^k=k^jk^ky2+Δ~jk12(1y2),02πdϕ2πp^ip^jp^k=k^ik^jk^ky3+(k^iΔ~jk+k^jΔ~ki+k^kΔ~ij)12y(1y2),02πdϕ2πp^ip^jp^kp^l=k^ik^jk^kk^ly4+(Δ~ijk^kk^l+Δ~jkk^lk^i+Δ~klk^ik^j+Δ~ilk^jk^k+Δ~ikk^jk^l+Δ~jlk^ik^k)×12y2(1y2)+(Δ~ijΔ~kl+Δ~ikΔ~jl+Δ~ilΔ~jk)18(1y2)2,\begin{split}\int_{0}^{2\pi}\frac{d\phi}{2\pi}{\hat{p}}^{k}&={\hat{k}}^{k}y\,,\\ \int_{0}^{2\pi}\frac{d\phi}{2\pi}{\hat{p}}^{j}{\hat{p}}^{k}&={\hat{k}}^{j}{\hat{k}}^{k}y^{2}+\tilde{\Delta}^{jk}\frac{1}{2}(1-y^{2})\,,\\ \int_{0}^{2\pi}\frac{d\phi}{2\pi}{\hat{p}}^{i}{\hat{p}}^{j}{\hat{p}}^{k}&={\hat{k}}^{i}{\hat{k}}^{j}{\hat{k}}^{k}y^{3}+\Bigl{(}{\hat{k}}^{i}\tilde{\Delta}^{jk}+{\hat{k}}^{j}\tilde{\Delta}^{ki}+{\hat{k}}^{k}\tilde{\Delta}^{ij}\Bigr{)}\frac{1}{2}y(1-y^{2})\,,\\ \int_{0}^{2\pi}\frac{d\phi}{2\pi}{\hat{p}}^{i}{\hat{p}}^{j}{\hat{p}}^{k}{\hat{p}}^{l}&={\hat{k}}^{i}{\hat{k}}^{j}{\hat{k}}^{k}{\hat{k}}^{l}y^{4}\\ &\quad+\Bigl{(}\tilde{\Delta}^{ij}{\hat{k}}^{k}{\hat{k}}^{l}+\tilde{\Delta}^{jk}{\hat{k}}^{l}{\hat{k}}^{i}+\tilde{\Delta}^{kl}{\hat{k}}^{i}{\hat{k}}^{j}+\tilde{\Delta}^{il}{\hat{k}}^{j}{\hat{k}}^{k}+\tilde{\Delta}^{ik}{\hat{k}}^{j}{\hat{k}}^{l}+\tilde{\Delta}^{jl}{\hat{k}}^{i}{\hat{k}}^{k}\Bigr{)}\\ &\qquad\times\frac{1}{2}y^{2}(1-y^{2})\\ &\quad+\Bigl{(}\tilde{\Delta}^{ij}\tilde{\Delta}^{kl}+\tilde{\Delta}^{ik}\tilde{\Delta}^{jl}+\tilde{\Delta}^{il}\tilde{\Delta}^{jk}\Bigr{)}\frac{1}{8}(1-y^{2})^{2}\,,\end{split} (209)

where we introduce Δ~ij=δijk^ik^j\tilde{\Delta}_{ij}=\delta_{ij}-{\hat{k}}_{i}{\hat{k}}_{j} (we note Δμν=ξμξνημν\Delta_{\mu\nu}=\xi_{\mu}\xi_{\nu}-\eta_{\mu\nu}). We prepare the following integrals:

x211𝑑yynxy={x(I1)(n=1)x2(I1)(n=2)x3+x3(I1)(n=3)x23+x4(I1)(n=4).\displaystyle\frac{x}{2}\int_{-1}^{1}dy\frac{y^{n}}{x-y}=\begin{cases}\displaystyle\,x(I-1)\quad&(n=1)\vspace{1em}\\ \displaystyle\,x^{2}(I-1)\quad&(n=2)\vspace{1em}\\ \displaystyle-\frac{x}{3}+x^{3}(I-1)\quad&(n=3)\vspace{0.5em}\\ \displaystyle-\frac{x^{2}}{3}+x^{4}(I-1)\quad&(n=4)\end{cases}\quad. (210)

These yield

Ij(x)\displaystyle I^{j}(x) =\displaystyle= k^jx(I1),\displaystyle{\hat{k}}^{j}x(I-1)\,, (211)
Ijk(x)\displaystyle I^{jk}(x) =\displaystyle= k^jk^kx2(I1)+12Δ~jk(Ix2(I1)),\displaystyle{\hat{k}}^{j}{\hat{k}}^{k}x^{2}(I-1)+\frac{1}{2}\tilde{\Delta}^{jk}\biggl{(}I-x^{2}(I-1)\biggr{)}\,, (212)
Iijk(x)\displaystyle I^{ijk}(x) =\displaystyle= k^ik^jk^k(x3+x3(I1))\displaystyle{\hat{k}}^{i}{\hat{k}}^{j}{\hat{k}}^{k}\Bigl{(}-\frac{x}{3}+x^{3}(I-1)\Bigr{)} (213)
+12(k^iΔ~jk+k^jΔ~ki+k^kΔ~ij)(x(I1)+x3x3(I1)),\displaystyle+\frac{1}{2}\Bigl{(}{\hat{k}}^{i}\tilde{\Delta}^{jk}+{\hat{k}}^{j}\tilde{\Delta}^{ki}+{\hat{k}}^{k}\tilde{\Delta}^{ij}\Bigr{)}\Bigl{(}x(I-1)+\frac{x}{3}-x^{3}(I-1)\Bigr{)}\,,
Iijkl(x)\displaystyle I^{ijkl}(x) =\displaystyle= (x23+x4(I1))k^ik^jk^kk^l\displaystyle\biggl{(}-\frac{x^{2}}{3}+x^{4}(I-1)\biggr{)}{\hat{k}}^{i}{\hat{k}}^{j}{\hat{k}}^{k}{\hat{k}}^{l} (214)
+12(x2(I1)+x23x4(I1))\displaystyle+\frac{1}{2}\biggl{(}x^{2}(I-1)+\frac{x^{2}}{3}-x^{4}(I-1)\biggr{)}
×(Δ~ijk^kk^l+Δ~jkk^lk^i+Δ~klk^ik^j+Δ~ilk^jk^k+Δ~ikk^jk^l+Δ~jlk^ik^k)\displaystyle\qquad\times\Bigl{(}\tilde{\Delta}^{ij}{\hat{k}}^{k}{\hat{k}}^{l}+\tilde{\Delta}^{jk}{\hat{k}}^{l}{\hat{k}}^{i}+\tilde{\Delta}^{kl}{\hat{k}}^{i}{\hat{k}}^{j}+\tilde{\Delta}^{il}{\hat{k}}^{j}{\hat{k}}^{k}+\tilde{\Delta}^{ik}{\hat{k}}^{j}{\hat{k}}^{l}+\tilde{\Delta}^{jl}{\hat{k}}^{i}{\hat{k}}^{k}\Bigr{)}
+18(I2x2(I1)x23+x4(I1))(Δ~ijΔ~kl+Δ~jkΔ~li+Δ~ilΔ~jk).\displaystyle+\frac{1}{8}\biggl{(}I-2x^{2}(I-1)-\frac{x^{2}}{3}+x^{4}(I-1)\biggr{)}\Bigl{(}\tilde{\Delta}^{ij}\tilde{\Delta}^{kl}+\tilde{\Delta}^{jk}\tilde{\Delta}^{li}+\tilde{\Delta}^{il}\tilde{\Delta}^{jk}\Bigr{)}\,.\quad

In particular, the asymptotic forms of Ij1jnI_{j_{1}\cdots j_{n}} in the dynamical limit x1x\gg 1 are

I\displaystyle I \displaystyle\simeq 1+13x2+15x4+17x6+O(x8),\displaystyle 1+\frac{1}{3x^{2}}+\frac{1}{5x^{4}}+\frac{1}{7x^{6}}+O(x^{-8})\,, (215)
Ij\displaystyle I^{j} \displaystyle\simeq k^j3x+O(x3),\displaystyle\frac{{\hat{k}}^{j}}{3x}+O(x^{-3})\,, (216)
Ijk\displaystyle I^{jk} \displaystyle\simeq (13+115x2)δjk+215x2k^jk^k+O(x4),\displaystyle\biggl{(}\frac{1}{3}+\frac{1}{15x^{2}}\biggr{)}\delta^{jk}+\frac{2}{15x^{2}}{\hat{k}}^{j}{\hat{k}}^{k}+O(x^{-4})\,, (217)
Iijk\displaystyle I^{ijk} \displaystyle\simeq 115x(k^iδjk+k^jδki+k^kδij)+O(x3),\displaystyle\frac{1}{15x}\Bigl{(}{\hat{k}}^{i}\delta^{jk}+{\hat{k}}^{j}\delta^{ki}+{\hat{k}}^{k}\delta^{ij}\Bigr{)}+O(x^{-3})\,, (218)
Iijkl\displaystyle I^{ijkl} \displaystyle\simeq (115+1105x2)(δijδkl+δikδil+δilδjk)\displaystyle\biggl{(}\frac{1}{15}+\frac{1}{105x^{2}}\biggr{)}\Bigl{(}\delta^{ij}\delta^{kl}+\delta^{ik}\delta^{il}+\delta^{il}\delta^{jk}\Bigr{)} (219)
+2105x2(δijk^kk^l+δjkk^lk^i+δklk^ik^j+δilk^jk^k+δikk^jk^l+δjlk^ik^k)+O(x4).\displaystyle+\frac{2}{105x^{2}}\Bigl{(}\delta^{ij}{\hat{k}}^{k}{\hat{k}}^{l}+\delta^{jk}{\hat{k}}^{l}{\hat{k}}^{i}+\delta^{kl}{\hat{k}}^{i}{\hat{k}}^{j}+\delta^{il}{\hat{k}}^{j}{\hat{k}}^{k}+\delta^{ik}{\hat{k}}^{j}{\hat{k}}^{l}+\delta^{jl}{\hat{k}}^{i}{\hat{k}}^{k}\Bigr{)}+O(x^{-4}).\qquad\quad

On the other hand, in the static limit x1x\ll 1, we find Ij1jnO(x)I_{j_{1}\cdots j_{n}}\simeq O(x).

References