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

Chiral anomaly in non-relativistic systems: Berry curvature and chiral kinetic theory

Lan-Lan Gao Physics Department and Center for Particle Physics and Field Theory, Fudan University, Shanghai 200438, China    Xu-Guang Huang [email protected] Physics Department and Center for Particle Physics and Field Theory, Fudan University, Shanghai 200438, China Key Laboratory of Nuclear Physics and Ion-beam Application (MOE), Fudan University, Shanghai 200433, China
Abstract

Chiral anomaly and the novel quantum phenomena it induces have been widely studied for Dirac and Weyl fermions. In most typical cases, the Lorentz covariance is assumed and thus the linear dispersion relations are maintained. However, in realistic materials, such as Dirac and Weyl semimetals, the non-linear dispersion relations appear naturally. We develop a kinetic framework to study the chiral anomaly for Weyl fermions with non-linear dispersions by using the methods of Wigner function and semi-classical equations of motion. In this framework, the chiral anomaly is sourced by Berry monopoles in momentum space and could be enhanced or suppressed due to the windings around the Berry monopoles. Our results can help understand the chiral anomaly-induced transport phenomena in non-relativistic systems.

pacs:
11.30.Rd, 72.10.Bg

I Introduction

During the development of condensed matter physics, the concepts and theories proficient in describing relativistic particles have been often found very useful. Among them, in recent years, the chiral anomaly which was first encountered in particle physics is particularly eye-catching in the studies of three-dimensional topological materials such as Weyl and Dirac semimetals (Yan and Felser, 2017; Armitage et al., 2018; Lv et al., 2021). In Weyl semimetals, the conduction and valence bands touch at some points (called Weyl nodes) in the Brillouin zone. The Weyl nodes always appear in pairs with opposite chirality around which the low-energy excitations have linear dispersion relations and can be described by Weyl Hamiltonian. In Dirac semimetals, the pairs of Weyl nodes of opposite chirality overlap to form Dirac nodes. The Weyl and Dirac semimetals exhibit a number of novel transport phenomena such as the chiral magnetic effect (CME) Kharzeev et al. (2008); Fukushima et al. (2008) which is a consequence of the chiral anomaly and induces special magnetoresistance Li et al. (2016); Huang et al. (2015); Xiong et al. (2015); Zhang et al. (2016), and novel modification of Casimir effect Fukushima et al. (2019); Rong et al. (2021). Interestingly, the CME has also been widely discussed in particle and nuclear physics and is under intensive search in relativistic heavy-ion collision experiments Kharzeev et al. (2016); Huang (2016a); Hattori and Huang (2017); Kharzeev and Liao (2021).

Although the Weyl semimetals are considered to be described by Weyl fermions, the realistic excitations in solids can be different from the ideal Weyl Hamiltonian 𝝈𝒑\propto\bm{\sigma}\cdot{\bm{p}} with 𝝈\bm{\sigma} the Pauli matrices and 𝒑{\bm{p}} the momentum away from the Weyl node. Their dispersion relations typically have tiltings, warpings, and non-linear momentum dependence, which are absent in the ideal Weyl Hamiltonian. These new features may be modeled by an extended Weyl Hamiltonian Fang et al. (2012); Soluyanov et al. (2015); Sun et al. (2015); Huang et al. (2017)

H=𝑲(𝒑)𝝈+K0(𝒑)H=\bm{K}({\bm{p}})\cdot\bm{\sigma}+K_{0}({\bm{p}}) (1)

with K0K_{0} and 𝑲\bm{K} polynomials of 𝒑{\bm{p}}. A number of interesting phenomena due to the tiltings, warpings, and non-linearity of the Weyl nodes are discussed including, e.g., the modified anomalous Hall effect Zyuzin and Tiwari (2016); Nandy et al. (2017) and the enhanced CME Sharma et al. (2017); Yu et al. (2016); Wei et al. (2018).

Moreover, generalizing the ideal Weyl Hamiltonian to include non-linear momentum dependence is also natural from the view point of low-energy effective theory. The band structure of a real material is usual very complex. Even for the usual Weyl semimetals, the linear Weyl Hamiltonian is considered as a lowest-order effective description valid only near the Weyl nodes. To describe physics away from the Weyl nodes, terms quadratic or even higher-power in momentum need to be included.

In this paper, we will first focus on a simple example of the Hamiltonian (1), given by

H=λ𝝈𝒑+α𝒑2,H=\lambda\bm{\sigma}\cdot\bm{p}+\alpha\bm{p}^{2}, (2)

where λ=±\lambda=\pm denotes the chirality of the Weyl node and the parameter α\alpha depends on the band structure of the material which is assumed to be positive to guarantee the stability of the system. Our purpose is to develop a chiral kinetic theory (CKT) from quantum field theory to describe the semi-classical physics of Hamiltonian (2) and to reveal how the chiral anomaly is self-consistently encoded in such a framework. The underlying methodology is parallel to its relativistic counterpart for relativistic Dirac or Weyl fermions that has been intensively discussed over the past few years Gao et al. (2012); Chen et al. (2013); Huang et al. (2018); Gao et al. (2018); Hidaka et al. (2017); Carignano et al. (2018); Liu et al. (2019); Sheng et al. (2020); Huang et al. (2020); Hattori et al. (2021); Chen and Lin (2021). We then will extend the discussion to the more general Hamiltonian (1).

The CKT for relativistic systems consists of a set of Lorentz covariant semi-classical kinetic equations that can well describe the anomalous transport effects like the CME and chiral vortical effect (CVE) Liu et al. (2019); Stephanov and Yin (2012); Chen et al. (2014); Huang and Sadofyev (2019). It also provides a convenient tool to explore the topological characteristics of the Dirac or Weyl fermions through the effects of Berry curvature in phase space. For example, the effects of external magnetic or gravitational field and the transports of Dirac fermions in curved space-time can be well expressed through the chiral kinetic theory Liu et al. (2019); Chen et al. (2014); Huang and Sadofyev (2019). The relativistic CKT has also been applied to condensed matter systems to study Berry-curvature induced transport phenomena Gorbar et al. (2017); Landsteiner (2014); Gao et al. (2021).

This paper is organized as follows. In section II , We will derive the CKT for non-relativistic Weyl fermions of type (2) by using the Wigner function method. We will show that, at the \hbar order (corresponding to one-loop approximation in field-theory calculation), the divergence of the chiral current receives a contribution from the singularity in phase space residing exactly at the Weyl node and gives the chiral anomaly in the same form as derived from relativistic field theory. In section III, we will directly derive the semi-classical equations of motion for Weyl fermions of type (1) and then check the anomaly relation. In section IV, we give comments on our results and discuss potential future application and extension of our results. Throughout this paper, we choose the Fermi velocity vF=1v_{F}=1 and the coupling constant e=1e=1. To simplify the notations, we will use the four-vector representation like xμ=(x0=t,𝒙)x^{\mu}=(x_{0}=t,{\bm{x}}) with index μ\mu from 0 to 33, pμ=(p0,𝒑)p^{\mu}=(p_{0},{\bm{p}}), px=p0t𝒑𝒙p\cdot x=p_{0}t-{\bm{p}}\cdot{\bm{x}}, and use ημν=ημν=diag(1,1,1,1)\eta^{\mu\nu}=\eta_{\mu\nu}={\rm diag}(1,-1,-1,-1) to raise and lower indices though no Lorentz invariance is assumed.

II Wigner function and chiral anomaly

The Wigner function W(x,p)W(x,p) is a generalization of the classical phase-space distribution function to quantum physics. It is defined as the ensemble average of Wigner operator given by

W^(x,p)=d4yeipy/ψ^(x)eyD/2eyD/2ψ^(x),\begin{split}\hat{W}(x,p)=\int d^{4}ye^{-ip\cdot y/\hbar}{\hat{\psi}}^{\dagger}(x)e^{y\cdot\overleftarrow{D}^{*}/2}\otimes e^{-y\cdot D/2}\hat{\psi}(x),\end{split} (3)

where ψ^(x)\hat{\psi}(x) is a Weyl spinor operator and [ψ^ψ^]ab=ψ^bψ^a[\hat{\psi}^{\dagger}\otimes\hat{\psi}]_{ab}=\hat{\psi}^{\dagger}_{b}\hat{\psi}_{a} with a,b=1a,b=1-22 being the spinor indices. Note that we have considered the presence of an external electromagnetic field AμA_{\mu} so that DμD_{\mu} is the covariant derivative acting on the Weyl spinor:

Dμψ^(x)=(μ+iAμ/)ψ^(x).D_{\mu}\hat{\psi}(x)=\Bigl{(}\partial_{\mu}+iA_{\mu}/\hbar\Bigr{)}\hat{\psi}(x)\,. (4)

The Wigner function is given by W(x,p)=W^(x,p)W(x,p)=\langle\hat{W}(x,p)\rangle where \langle\cdots\rangle denotes the ensemble average. The equations of motion (Heisenberg equations) of ψ^\hat{\psi} and ψ^\hat{\psi}^{\dagger} for Hamiltonian (2) are given by:

iDtψ^+α𝑫2ψ^+iλ𝝈𝑫ψ^=0,\displaystyle iD_{t}\hat{\psi}+\hbar\alpha{\bm{D}}^{2}\hat{\psi}+i\lambda{\bm{\sigma}}\cdot{\bm{D}}\hat{\psi}=0, (5)
iDtψ^α𝑫2ψ^+iλ𝑫ψ^𝝈=0.\displaystyle iD^{*}_{t}\hat{\psi}^{\dagger}-\hbar\alpha{\bm{D}}^{*2}\hat{\psi}^{\dagger}+i\lambda{\bm{D}}^{*}\hat{\psi}^{\dagger}\cdot{\bm{\sigma}}=0. (6)

We will consider \hbar as being much smaller than the classical action of the system and thus make an expansion in \hbar. Combining Eqs. (5)-(6) and the definition (3) for Wigner operator, we derive the equation of motion for Wigner function accurate at O(2)O(\hbar^{2}) as

[i2Δt+p0+λσi(i2Δi+pi)α(i2Δi+pi)(i2Δi+pi)]W=0,\begin{split}&\bigg{[}\frac{i\hbar}{2}\Delta_{t}+p_{0}+\lambda\sigma^{i}\left(\frac{i\hbar}{2}\Delta_{i}+p_{i}\right)\\ &\quad\quad\quad-\alpha\left(\frac{i\hbar}{2}\Delta_{i}+p_{i}\right)\left(\frac{i\hbar}{2}\Delta_{i}+p_{i}\right)\bigg{]}W=0,\end{split} (7)

where Δμ=μFμνpν\Delta_{\mu}=\partial_{\mu}-F_{\mu\nu}\partial_{p}^{\nu} (Here, Δ0=Δt\Delta_{0}=\Delta_{t}) with Fμν=μAννAμF_{\mu\nu}=\partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu} the field strength, pμ=/pμ\partial^{\mu}_{p}=\partial/\partial p_{\mu}, and repeated indices are summed. In deriving this equation, we have made an expansion in \hbar with the power counting rule: pμO(1),AμO(1)p^{\mu}\sim O(1),A_{\mu}\sim O(1), and yμipμO()y^{\mu}\sim i\hbar\partial^{\mu}_{p}\sim O(\hbar), and have used the following equations which are accurate up to O()O(\hbar):

[Dμ,eyD2]=iyν2(Fμνyρ4ρFμν)eyD2+O(2),\displaystyle[D_{\mu},e^{-\frac{y\cdot D}{2}}]=-\frac{iy^{\nu}}{2\hbar}\left(F_{\mu\nu}-\frac{y^{\rho}}{4}\partial_{\rho}F_{\mu\nu}\right)e^{-\frac{y\cdot D}{2}}+O(\hbar^{2})\,,
[μy,eyD2]=12eyD2(Dμiyν4Fμν+iyνyρ12ρFμν)\displaystyle[\partial_{\mu}^{y},e^{-\frac{y\cdot D}{2}}]=-\frac{1}{2}e^{-\frac{y\cdot D}{2}}\left(D_{\mu}-\frac{iy^{\nu}}{4\hbar}F_{\mu\nu}+\frac{iy^{\nu}y^{\rho}}{12\hbar}\partial_{\rho}F_{\mu\nu}\right)
+O(2).\displaystyle\mkern-120.0mu+O(\hbar^{2})\,.

To reveal the physical content in Eq. (LABEL:eq:Weq-full), we expand the Wigner function in su(2)su(2) algebra, W=12(𝒱0+λσi𝒱i)W=\frac{1}{2}(\mathcal{V}_{0}+\lambda\sigma_{i}\mathcal{V}_{i}). Note that the Wigner function is Hermitian so that the coefficients 𝒱0\mathcal{V}_{0} and 𝒱i\mathcal{V}_{i} are real. The physical meanings of 𝒱0\mathcal{V}_{0} and 𝒱i\mathcal{V}_{i} become clear if we express them as 𝒱0(x,p)=TrW(x,p)\mathcal{V}_{0}(x,p)={\rm Tr}\,W(x,p) and 𝓥(x,p)=λTr[𝝈W(x,p)]\bm{\mathcal{V}}(x,p)=\lambda{\rm Tr}\,[\bm{\sigma}W(x,p)]. After integration over momentum, 𝒱0(x)=d4p(2π)4𝒱0(x,p)=ψ^(x)ψ^(x)\mathcal{V}_{0}(x)=\int\frac{d^{4}p}{(2\pi)^{4}}\mathcal{V}_{0}(x,p)=\langle\hat{\psi}^{\dagger}(x)\hat{\psi}(x)\rangle 111In principle, the momentum-space measure should be d4p/(2π)4d^{4}p/(2\pi\hbar)^{4}, but the \hbar here does not influence any of our discussions, so we omit it. and 𝓥(x)=λψ^(x)𝝈ψ^(x)\bm{\mathcal{V}}(x)=\lambda\langle\hat{\psi}^{\dagger}(x)\bm{\sigma}\hat{\psi}(x)\rangle. Thus 𝒱0(x,p)\mathcal{V}_{0}(x,p) is considered as the phase-space particle number distribution and 𝓥(x,p)\bm{\mathcal{V}}(x,p) can be considered as (part of) the particle current (See below) or spin distribution in phase space. These coefficients satisfy the following four equations which can be easily extracted from Eq. (LABEL:eq:Weq-full):

(p0α𝒑2+α24𝚫2)𝒱ipi𝒱0λ2ϵijkΔj𝒱k=0,\displaystyle\left(p_{0}-\alpha{\bm{p}}^{2}+\alpha\frac{\hbar^{2}}{4}\bm{\Delta}^{2}\right)\mathcal{V}^{i}-p^{i}\mathcal{V}_{0}-\lambda\frac{\hbar}{2}\epsilon^{ijk}\Delta_{j}\mathcal{V}^{k}=0,\;\;\;\;\;\;\; (8)
2(Δt𝒱i2αpjΔj𝒱i+Δi𝒱0)+λϵijkpj𝒱k=0,\displaystyle\frac{\hbar}{2}\left(\Delta_{t}\mathcal{V}^{i}-2\alpha p_{j}\Delta_{j}\mathcal{V}^{i}+\Delta_{i}\mathcal{V}_{0}\right)+\lambda\epsilon^{ijk}p_{j}\mathcal{V}^{k}=0, (9)
(p0α𝒑2+α24𝚫2)𝒱0+pi𝒱i=0,\displaystyle\left(p_{0}-\alpha{\bm{p}}^{2}+\alpha\frac{\hbar^{2}}{4}\bm{\Delta}^{2}\right)\mathcal{V}_{0}+p_{i}\mathcal{V}^{i}=0, (10)
(Δt2αpjΔj)𝒱0+Δi𝒱i=0.\displaystyle(\Delta_{t}-2\alpha p_{j}\Delta_{j})\mathcal{V}_{0}+\Delta_{i}\mathcal{V}^{i}=0. (11)

We now solve Eqs. (8)-(11) order by order in \hbar. At zeroth order (namely, the classical level), combining Eqs. (8)-(10), one easily finds that

𝒱0(0)\displaystyle\mathcal{V}^{(0)}_{0} =\displaystyle= 4πp~0f(0)(x,p)δ(p~2),\displaystyle 4\pi\tilde{p}_{0}f^{(0)}(x,p)\delta(\tilde{p}^{2}), (12)
𝒱i(0)\displaystyle\mathcal{V}^{(0)}_{i} =\displaystyle= 4πp~if(0)(x,p)δ(p~2),\displaystyle 4\pi\tilde{p}_{i}f^{(0)}(x,p)\delta(\tilde{p}^{2}), (13)

where we have introduced the shorthand notation p~μ=(p~0,𝒑~)\tilde{p}^{\mu}=(\tilde{p}^{0},\tilde{{\bm{p}}}) with p~0=p~0=p0α𝒑2\tilde{p}^{0}=\tilde{p}_{0}=p_{0}-\alpha{\bm{p}}^{2}, p~i=p~i=pi\tilde{p}^{i}=-\tilde{p}_{i}=p^{i}, and p~2=p~02𝒑~2\tilde{p}^{2}=\tilde{p}^{2}_{0}-\tilde{{\bm{p}}}^{2}. The function f(0)f^{(0)} represents the classical distribution function in phase space Elze et al. (1986); Vasak et al. (1987). Substituting 𝒱0(0)\mathcal{V}^{(0)}_{0} and 𝒱i(0)\mathcal{V}^{(0)}_{i} into Eq. (11), we obtain the classical (collisionless) kinetic equation for f(0)f^{(0)}:

δ(p~2)[(p0α𝒑2)(Δt+2α𝒑𝚫)+𝒑𝚫]f(0)=0.\delta(\tilde{p}^{2})\left[(p_{0}-\alpha{\bm{p}}^{2})(\Delta_{t}+2\alpha{\bm{p}}\cdot\bm{\Delta})+{\bm{p}}\cdot\bm{\Delta}\right]f^{(0)}=0. (14)

To proceed to O()O(\hbar) order, it is convenient to re-write Eqs. (8)-(11) at O()O(\hbar) order in a formally covariant form:

p~μ𝒱νp~ν𝒱μ+λ2ϵμνρσΔ~ρ𝒱σ=0,\displaystyle\tilde{p}_{\mu}\mathcal{V}_{\nu}-\tilde{p}_{\nu}\mathcal{V}_{\mu}+\lambda\frac{\hbar}{2}\epsilon_{\mu\nu\rho\sigma}\tilde{\Delta}^{\rho}\mathcal{V}^{\sigma}=0, (15)
p~𝒱=0,\displaystyle\tilde{p}\cdot\mathcal{V}=0, (16)
Δ~𝒱=0,\displaystyle\tilde{\Delta}\cdot\mathcal{V}=0, (17)

where the four-vectors are defined by Δ~μ=(Δt2αpjΔj,Δi)\tilde{\Delta}_{\mu}=(\Delta_{t}-2\alpha p_{j}\Delta_{j},\Delta_{i}) and 𝒱μ=(𝒱0,𝒱i)\mathcal{V}^{\mu}=({\mathcal{V}_{0},\mathcal{V}^{i}}), and the four-dimensional Levi-Civita symbol is normalized to ϵ0ijk=ϵ0ijk=ϵijk\epsilon^{0ijk}=-\epsilon_{0ijk}=\epsilon^{ijk}. The contraction is performed with metric ημν\eta_{\mu\nu}, e.g., p~𝒱=ημνp~μ𝒱ν=p~0𝒱0𝒑~𝓥\tilde{p}\cdot\mathcal{V}=\eta_{\mu\nu}\tilde{p}^{\mu}\mathcal{V}^{\nu}=\tilde{p}_{0}\mathcal{V}_{0}-\tilde{\bm{p}}\cdot\bm{\mathcal{V}}. This set of equations are in the same form as their counterparts in relativistic CKT and thus can be solved in the same way as stressed in Refs. Hidaka et al. (2017); Liu et al. (2019); Huang et al. (2018). Multiplying Eq. (15) by an arbitrary vector nμn^{\mu} satisfying nμnμ=1n^{\mu}n_{\mu}=1 and np~0n\cdot\tilde{p}\neq 0 and combining it with the zeroth order results given in Eqs. (12)-(13) and the constraint (16), we finally obtain

𝒱μ=4πδ(p~2)[p~μλp~2G~μνp~ν+λΣμνΔ~ν]f+O(2),\displaystyle\mathcal{V}^{\mu}=4\pi\delta(\tilde{p}^{2})\left[\tilde{p}^{\mu}-\lambda\frac{\hbar}{\tilde{p}^{2}}\tilde{G}^{\mu\nu}\tilde{p}_{\nu}+\lambda\hbar\Sigma^{\mu\nu}\tilde{\Delta}_{\nu}\right]f+O(\hbar^{2}),

where Σμν=ϵμνρσp~ρnσ/(2p~n)\Sigma_{\mu\nu}=\epsilon_{\mu\nu\rho\sigma}\tilde{p}^{\rho}n^{\sigma}/(2\tilde{p}\cdot n) is called the spin tensor in a frame specified by vector nμn^{\mu}, f=f(0)+f(1)f=f^{(0)}+f^{(1)} is the distribution function up to order O()O(\hbar), and G~μν=(1/2)ϵμνρσGρσ\tilde{G}^{\mu\nu}=(1/2)\epsilon^{\mu\nu\rho\sigma}G_{\rho\sigma} is the dual tensor of GμνG_{\mu\nu} which is defined by Gμν=Δ~μp~νG_{\mu\nu}=-\tilde{\Delta}_{\mu}\tilde{p}_{\nu} and related to the field strength tensor via G0i=Gi0=F0i2αpjFji=Ei+2αϵijkpjBkG_{0i}=-G_{i0}=F_{0i}-2\alpha p_{j}F_{ji}=E^{i}+2\alpha\epsilon^{ijk}p^{j}B^{k} and Gij=Fij=ϵijkBkG_{ij}=F_{ij}=-\epsilon^{ijk}B^{k} with 𝑬,𝑩\bm{E},\bm{B} the electric and magnetic fields. Substituting Eq. (II) into Eq. (17), we obtain the O()O(\hbar)-order kinetic equation for ff:

δ(p~2λΣαβGαβ)[p~Δ~+λ(nσp~nG~σνΔ~ν+Δ~μΣμνΔ~ν)]f(t,𝒙,p0,𝒑)=0.\begin{split}&\delta(\tilde{p}^{2}-\lambda\hbar\Sigma_{\alpha\beta}G^{\alpha\beta})\bigg{[}\tilde{p}\cdot\tilde{\Delta}\\ &\quad\quad+\lambda\hbar\left(\frac{n_{\sigma}}{\tilde{p}\cdot n}\tilde{G}^{\sigma\nu}\tilde{\Delta}_{\nu}+\tilde{\Delta}_{\mu}\Sigma^{\mu\nu}\tilde{\Delta}_{\nu}\right)\bigg{]}f(t,{\bm{x}},p_{0},{\bm{p}})=0.\end{split} (19)

To derive this equation, we have used the following relations: δ(x)=δ(x)/x\delta^{\prime}(x)=-\delta(x)/x, δ′′(x)=2δ(x)/x\delta^{\prime\prime}(x)=-2\delta^{\prime}(x)/x, Δ~μG~μν=0\tilde{\Delta}_{\mu}\tilde{G}^{\mu\nu}=0, and G~μρGμσp~ρp~σ=G~αβGαβp~2/4\tilde{G}^{\mu\rho}G_{\mu\sigma}\tilde{p}_{\rho}\tilde{p}^{\sigma}=\tilde{G}^{\alpha\beta}G_{\alpha\beta}\tilde{p}^{2}/4 which can be proven using the Schouten identity ϵ[μνρσp~λ]=0\epsilon^{[\mu\nu\rho\sigma}\tilde{p}^{\lambda]}=0 with [][\cdots] meaning the anti-symmetrization operation. The form of Eq. (19) shows clearly similarity with its relativistic counterpart in, e.g., Ref. Liu et al. (2019).

To see the physical content in Eq. (19) more clearly, it is instructive to choose nμ=(1,0,0,0)n^{\mu}=(1,0,0,0) (the laboratory frame). The Dirac δ\delta function in Eq. (19) gives the on-shell condition (dispersion relation) for the fermions in which the term λΣG-\lambda\hbar\Sigma\cdot G expresses a quantum correction due to the coupling between the magnetic moment and magnetic field:

p0=εp=±|𝒑|(1𝛀λ𝑩)+α𝒑2,\displaystyle p_{0}=\varepsilon_{p}=\pm|{\bm{p}}|(1\mp\hbar\bm{\Omega}_{\lambda}\cdot\bm{B})+\alpha{\bm{p}}^{2}, (20)

where the upper (lower) sign corresponds to conduction (valance) band of Hamiltonian (2). In the following, we focus on the upper sign. The magnetic moment is determined by a Berry curvature

𝛀λ=λ𝒑2|𝒑|3\displaystyle\bm{\Omega}_{\lambda}=\lambda\frac{{\bm{p}}}{2|{\bm{p}}|^{3}} (21)

of Berry monopole residing at 𝒑=𝟎{\bm{p}}=\bm{0} with charge λ\lambda (chirality): p𝛀λ=2πλδ(3)(𝒑)\bm{\nabla}_{p}\cdot\bm{\Omega}_{\lambda}=2\pi\lambda\delta^{(3)}({\bm{p}}). Our semi-classical scheme is meaningful only when the quantum correction does not break the band gap |𝒑||{\bm{p}}|; this requires |𝛀λ𝑩|1\hbar|\bm{\Omega}_{\lambda}\cdot\bm{B}|\ll 1 leading to |𝒑||𝑩||{\bm{p}}|\gg\sqrt{\hbar|\bm{B}|} (the adiabatic condition) at which our kinetic description applies. Integrating Eq. (19) over p~0\tilde{p}_{0} from 0 to \infty, we pickup the contribution of conduction fermions and, after a tedious calculation, find

{Gt+[𝒗p+(𝒗p𝛀λ)𝑩+𝑬~×𝛀λ]x+[𝑬~+𝒗p×𝑩+(𝑬~𝑩)𝛀λ]p}f(t,𝒙,𝒑)=0,\begin{split}&\Big{\{}\sqrt{G}\partial_{t}+\left[\bm{v}_{p}+\hbar(\bm{v}_{p}\cdot\bm{\Omega}_{\lambda})\bm{B}+\hbar\tilde{\bm{E}}\times\bm{\Omega}_{\lambda}\right]\cdot\bm{\nabla}_{x}\\ &\quad+\left[\tilde{\bm{E}}+\bm{v}_{p}\times\bm{B}+\hbar(\tilde{\bm{E}}\cdot\bm{B})\bm{\Omega}_{\lambda}\right]\cdot\bm{\nabla}_{p}\Big{\}}f(t,{\bm{x}},{\bm{p}})=0,\end{split} (22)

where f(t,𝒙,𝒑)=f(t,𝒙,p0=εp,𝒑)f(t,{\bm{x}},{\bm{p}})=f(t,{\bm{x}},p_{0}=\varepsilon_{p},{\bm{p}}), G=1+𝛀λ𝑩\sqrt{G}=1+\hbar\bm{\Omega}_{\lambda}\cdot\bm{B} is the quantum corrected measure of phase space (see below), 𝒗p=pεp\bm{v}_{p}=\bm{\nabla}_{p}\varepsilon_{p} is the single-particle velocity, and 𝑬~=𝑬xεp\tilde{\bm{E}}=\bm{E}-\bm{\nabla}_{x}\varepsilon_{p} is a modified electric field. To derive this equation, we have neglected all the O(2)O(\hbar^{2}) terms and used the relations (Note that μ\partial_{\mu} and pi\partial^{i}_{p} on the left-hand and right-hand sides act on different arguments.): μf(t,𝒙,𝒑)={[μ+(μεp)p0]f(t,𝒙,p0,𝒑)}|p0=εp\partial_{\mu}f(t,{\bm{x}},{\bm{p}})=\{[\partial_{\mu}+(\partial_{\mu}\varepsilon_{p})\partial^{0}_{p}]f(t,{\bm{x}},p_{0},{\bm{p}})\}|_{p_{0}=\varepsilon_{p}} and pif(t,𝒙,𝒑)={[pi+(piεp)p0]f(t,𝒙,p0,𝒑)}|p0=εp\partial^{i}_{p}f(t,{\bm{x}},{\bm{p}})=\{[\partial^{i}_{p}+(\partial^{i}_{p}\varepsilon_{p})\partial^{0}_{p}]f(t,{\bm{x}},p_{0},{\bm{p}})\}|_{p_{0}=\varepsilon_{p}}. The kinetic equation for valance fermions can be similarly obtained and the result is in the same form of Eq. (22) but with εp\varepsilon_{p} understood as being given by the lower sign in Eq. (20) and λ\lambda replaced by λ-\lambda in all the other terms.

By comparing Eq. (22) with the standard form (t+𝒙˙x+𝒑˙p)f=0(\partial_{t}+\dot{{\bm{x}}}\cdot\bm{\nabla}_{x}+\dot{{\bm{p}}}\cdot\bm{\nabla}_{p})f=0 of Boltzmann equation, we extract the single-particle equations of motion:

G𝒙˙\displaystyle\sqrt{G}\dot{{\bm{x}}} =\displaystyle= 𝒗p+(𝒗p𝛀λ)𝑩+𝑬~×𝛀λ,\displaystyle\bm{v}_{p}+\hbar(\bm{v}_{p}\cdot\bm{\Omega}_{\lambda})\bm{B}+\hbar\tilde{\bm{E}}\times\bm{\Omega}_{\lambda}, (23)
G𝒑˙\displaystyle\sqrt{G}\dot{{\bm{p}}} =\displaystyle= 𝑬~+𝒗p×𝑩+(𝑬~𝑩)𝛀λ.\displaystyle\tilde{\bm{E}}+\bm{v}_{p}\times\bm{B}+\hbar(\tilde{\bm{E}}\cdot\bm{B})\bm{\Omega}_{\lambda}. (24)

Similar set of equations of motion has been derived repeatedly in literature in different contexts (though some of them are not complete at \hbar order) for, e.g., band electrons in solids Sundaram and Niu (1999); Xiao et al. (2010, 2005); Son and Yamamoto (2012), trapped cold atoms Huang (2016b), neutrinos in supernovae Yamamoto (2016), quarks in quark-gluon plasma Stephanov and Yin (2012); Son and Yamamoto (2013); Chen et al. (2013), and photons Bliokh and Bliokh (2004); Onoda et al. (2004); Yamamoto (2017); Huang and Sadofyev (2019). A direct calculation leads to

tG+x(G𝒙˙)+p(G𝒑˙)=2πλδ(3)(𝒑)𝑬𝑩,\displaystyle\partial_{t}\sqrt{G}+\bm{\nabla}_{x}\cdot(\sqrt{G}\dot{{\bm{x}}})+\bm{\nabla}_{p}\cdot(\sqrt{G}\dot{{\bm{p}}})=2\pi\lambda\hbar\delta^{(3)}({\bm{p}})\bm{E}\cdot\bm{B},

where we have used the Maxwell equations x𝑩=0\bm{\nabla}_{x}\cdot\bm{B}=0 and t𝑩=x×𝑬\partial_{t}\bm{B}=-\bm{\nabla}_{x}\times\bm{E} and have omitted O(2)O(\hbar^{2}) terms. This identifies G\sqrt{G} as the invariant phase-space measure except for a singularity at 𝒑=𝟎{\bm{p}}=\bm{0} at which a Berry monopole resides. This Berry monopole contributes anomalous velocity which could lead to, for example, CME or anomalous Hall effect, and anomalous force, which could lead to chiral anomaly as we show now.

The U(1)U(1) Noether current jμ=(j0,ji)j^{\mu}=(j^{0},j^{i}) associated with Hamiltonian (2) is easily expressed by 𝒱μ\mathcal{V}^{\mu},

j0(x)\displaystyle j^{0}(x) =\displaystyle= ψ^(x)ψ^(x)=d4p(2π)4𝒱0(x,p),\displaystyle\langle\hat{\psi}^{\dagger}(x)\hat{\psi}(x)\rangle=\int\frac{d^{4}p}{(2\pi)^{4}}\mathcal{V}^{0}(x,p), (26)
ji(x)\displaystyle j^{i}(x) =\displaystyle= λψ^(x)σiψ^(x)iαψ^(x)(DiDi)ψ^(x)\displaystyle\lambda\langle\hat{\psi}^{\dagger}(x)\sigma^{i}\hat{\psi}(x)\rangle-i\hbar\alpha\langle\hat{\psi}^{\dagger}(x)(D_{i}-\overleftarrow{D}^{*}_{i})\hat{\psi}(x)\rangle (27)
=\displaystyle= d4p(2π)4[𝒱i(x,p)+2αpi𝒱0(x,p)].\displaystyle\int\frac{d^{4}p}{(2\pi)^{4}}\left[\mathcal{V}^{i}(x,p)+2\alpha p^{i}\mathcal{V}^{0}(x,p)\right].

Substituting Eq. (II) and after a lengthy but straightforward calculation, we obtain 222Here, we only present the contributions from conduction fermions. If the contributions of the valance fermions are taken into account, additional terms should be added to j0j^{0} and jij^{i} with the same forms as Eqs. (28) and (II) but with a minus sign for the magnetization current.

j0(x)\displaystyle j^{0}(x) =\displaystyle= d3𝒑(2π)3Gf(t,𝒙,𝒑),\displaystyle\int\frac{d^{3}{\bm{p}}}{(2\pi)^{3}}\sqrt{G}f(t,{\bm{x}},{\bm{p}}), (28)
𝒋(x)\displaystyle\bm{j}(x) =\displaystyle= d3𝒑(2π)3G[𝒙˙|𝒑|𝛀λ×x]f(t,𝒙,𝒑).\displaystyle\int\frac{d^{3}{\bm{p}}}{(2\pi)^{3}}\sqrt{G}\left[\dot{{\bm{x}}}-\hbar|{\bm{p}}|\bm{\Omega}_{\lambda}\times\bm{\nabla}_{x}\right]f(t,{\bm{x}},{\bm{p}}).

Note that the second term in Eq. (II) can be written as x×𝑴\bm{\nabla}_{x}\times\bm{M} with the magnetization 𝑴=G|𝒑|𝛀λfd3𝒑/(2π)3\bm{M}=\hbar\int\sqrt{G}|{\bm{p}}|\bm{\Omega}_{\lambda}fd^{3}{\bm{p}}/(2\pi)^{3} and thus represents a magnetization current. Now, using Eqs. (22) and (II), one can finds that the U(1)U(1) current is not conserved at O()O(\hbar):

μjμ=λ4π2𝑬𝑩f(t,𝒙,𝒑=𝟎).\partial_{\mu}j^{\mu}=\lambda\frac{\hbar}{4\pi^{2}}\bm{E}\cdot\bm{B}f(t,{\bm{x}},{\bm{p}}=\bm{0}). (30)

If the Fermi surface encloses the point 𝒑=𝟎{\bm{p}}=\bm{0} so that f(𝒑=𝟎)=1f({\bm{p}}=\bm{0})=1 at T=0T=0, the above relation recover exactly the chiral anomaly relation. Therefore, in the kinetic framework, the chiral anomaly is sourced by the Berry monopole at the Weyl node. Further more, although the presence of α𝒑2\alpha{\bm{p}}^{2} terms in Hamiltonian (2) modifies the single-particle dispersion relation and thus the kinetic equation, it does not change the form of chiral anomaly relation as long as the Fermi surface encloses the Berry monopole.

Before closing this section, we comment on the Wigner-function method in comparison with the methods in  Sundaram and Niu (1999); Xiao et al. (2010, 2005); Son and Yamamoto (2012); Huang (2016b) based on the wave-packet dynamics of band electrons. Unlike the wave-packet method which can be considered as a bottom-up treatment starting from the single-particle quantum mechanics of the band electrons, the starting point of the Wigner-function method is the underlying quantum field theory. It allows a top-down treatment of the quantum dynamics in a manner of an expansion in \hbar. The kinetic equation and the corresponding single-particle equations of motion can be systemcatically obtained to an arbitrary order in \hbar expansion. Moreover, the Wigner-function method can be easily extended to the study of bosons.

III Semi-classical equations of motion and Chiral anomaly

The above discusses focus on Hamiltonian (2), we now consider the more general Hamiltonian (1). We first derive the semi-classical equations of motion of single particle for Hamiltonian (1). This can be achieved by multiple methods. Here, we use the following well established result Stiepan and Teufel (2013):
Suppose H(𝒙,𝒌)=H0(𝒙,𝒌)+H1(𝒙,𝒌)+O(2)H({\bm{x}},{\bm{k}})=H_{0}({\bm{x}},{\bm{k}})+\hbar H_{1}({\bm{x}},{\bm{k}})+O(\hbar^{2}) with (𝒙,𝒌)({\bm{x}},{\bm{k}}) the canonical pairs of mechanical variables. For each isolated eigenvalue h0(𝒙,𝒌)h_{0}({\bm{x}},{\bm{k}}) of H0(𝒙,𝒌)H_{0}({\bm{x}},{\bm{k}}), there is associated a classical system with an \hbar-dependent Hamilton function h(𝒙,𝒌)=h0(𝒙,𝒌)+h1(𝒙,𝒌)+O(2)h({\bm{x}},{\bm{k}})=h_{0}({\bm{x}},{\bm{k}})+\hbar h_{1}({\bm{x}},{\bm{k}})+O(\hbar^{2}) and a modified symplectic form G(𝒙,𝒌)=G0+Ω(𝒙,𝒌)+O(2)G({\bm{x}},{\bm{k}})=G_{0}+\hbar\Omega({\bm{x}},{\bm{k}})+O(\hbar^{2}) (Ω\Omega is a generalized Berry curvature), where

h1\displaystyle h_{1} =\displaystyle= tr(H1π0)i2tr(π0{π0,H0h0}),\displaystyle{\rm tr}\left(H_{1}\pi_{0}\right)-\frac{i}{2}{\rm tr}\left(\pi_{0}\{\pi_{0},H_{0}-h_{0}\}\right), (31)
Ωαβ\displaystyle\Omega_{\alpha\beta} =\displaystyle= itr(π0[zαπ0,zβπ0]).\displaystyle-i{\rm tr}\left(\pi_{0}[\partial_{z^{\alpha}}\pi_{0},\partial_{z^{\beta}}\pi_{0}]\right). (32)

Here, the trace is over inner space (for our case, the spin) and 𝒙{\bm{x}} and 𝒌{\bm{k}} are collectively denoted by zαz^{\alpha} (α=1,,6\alpha=1,\cdots,6). π0\pi_{0} is spectral projection to the eigenenergy h0h_{0}, i.e., H0π0=h0π0H_{0}\pi_{0}=h_{0}\pi_{0} and π02=π0\pi_{0}^{2}=\pi_{0}. The zeroth order symplectic form is given by G0=iσy𝟏3G_{0}=i\sigma_{y}\otimes{\bf 1}_{3}. The Poisson bracket is defined by {A,B}=i=13(kiAxiBxiAkiB)\{A,B\}=\sum_{i=1}^{3}(\partial_{k^{i}}A\partial_{x^{i}}B-\partial_{x^{i}}A\partial_{k^{i}}B). The commutator is defined by [A,B]=ABBA[A,B]=AB-BA. If the Hamiltonian depends on time explicitly, H=H(t,𝒙,𝒌)H=H(t,{\bm{x}},{\bm{k}}), we just add the canonical pair (t,E)(t,E) and apply the previous procedure to a new Hamiltonian H~(t,𝒙,E,𝒌)=E+H(t,𝒙,𝒌)\tilde{H}(t,{\bm{x}},E,{\bm{k}})=E+H(t,{\bm{x}},{\bm{k}}). Its spectral projections π0(t,𝒙,𝒌)\pi_{0}(t,{\bm{x}},{\bm{k}}) are independent of EE and the classical Hamilton function is h(t,𝒙,E,𝒌)=E+h0(t,𝒙,𝒌)+h1(t,𝒙,𝒌)+O(2)h(t,{\bm{x}},E,{\bm{k}})=E+h_{0}(t,{\bm{x}},{\bm{k}})+\hbar h_{1}(t,{\bm{x}},{\bm{k}})+O(\hbar^{2}) with symplectic form G=G0+Ω(t,𝒙,𝒑)G=G_{0}+\hbar\Omega(t,{\bm{x}},{\bm{p}}), where hh and Ω\Omega are computed from the instantaneous Hamilton function H(t,𝒙,𝒑)H(t,{\bm{x}},{\bm{p}}) as before. Note that with symplectic form GG the semi-classical equations of motion are

Gαβz˙β=hzα,\displaystyle G_{\alpha\beta}\dot{z}^{\beta}=-\frac{\partial h}{\partial{z^{\alpha}}}, (33)

and the invariant phase-space measure is changed to

Gdet(Gαβ)=1+2(ΩxikiΩkixi)+O(2).\displaystyle\sqrt{G}\equiv\sqrt{{\rm det}\;(G_{\alpha\beta})}=1+\frac{\hbar}{2}(\Omega_{x^{i}k^{i}}-\Omega_{k^{i}x^{i}})+O(\hbar^{2}).

For Hamiltonian (1), H1=0H_{1}=0, and 𝒑=𝒌𝑨(t,𝒙){\bm{p}}={\bm{k}}-\bm{A}(t,{\bm{x}}) is the kinetic momentum if external electromagnetic field is applied. The eigenenergy and the corresponding projection are

h0\displaystyle h_{0} =\displaystyle= ±|𝑲(t,𝒙,𝒌)|+K0(t,𝒙,𝒌)+A0(t,𝒙),\displaystyle\pm|\bm{K}(t,{\bm{x}},{\bm{k}})|+K_{0}(t,{\bm{x}},{\bm{k}})+A_{0}(t,{\bm{x}}), (35)
π0\displaystyle\pi_{0} =\displaystyle= 12(1±𝑲^𝝈),\displaystyle\frac{1}{2}(1\pm\hat{\bm{K}}\cdot\bm{\sigma}), (36)

where 𝑲^=𝑲/|𝑲|\hat{\bm{K}}=\bm{K}/|\bm{K}|. Substituting them into Eqs. (31) and (32), we find

h1\displaystyle h_{1} =\displaystyle= 12𝑲^ki×𝑲^xi𝑲,\displaystyle\frac{1}{2}\frac{\partial\hat{\bm{K}}}{\partial k^{i}}\times\frac{\partial\hat{\bm{K}}}{\partial x^{i}}\cdot\bm{K}, (37)
Ωαβ\displaystyle\Omega_{\alpha\beta} =\displaystyle= ±12𝑲^zα×𝑲^zβ𝑲^,\displaystyle\pm\frac{1}{2}\frac{\partial\hat{\bm{K}}}{\partial z^{\alpha}}\times\frac{\partial\hat{\bm{K}}}{\partial z^{\beta}}\cdot\hat{\bm{K}}, (38)

and thus the semi-classical equations of motion read

x˙i\displaystyle\dot{x}^{i} =\displaystyle= (δij+Ωkixj)hkjΩkikjhxj+Ωkit,\displaystyle\left(\delta^{ij}+\hbar\Omega_{k^{i}x^{j}}\right)\frac{\partial h}{\partial k^{j}}-\hbar\Omega_{k^{i}k^{j}}\frac{\partial h}{\partial x^{j}}+\hbar\Omega_{k^{i}t}, (39)
k˙i\displaystyle\dot{k}^{i} =\displaystyle= (δij+Ωxikj)hxjΩxixjhkjΩxit.\displaystyle\left(-\delta^{ij}+\hbar\Omega_{x^{i}k^{j}}\right)\frac{\partial h}{\partial x^{j}}-\hbar\Omega_{x^{i}x^{j}}\frac{\partial h}{\partial k^{j}}-\hbar\Omega_{x^{i}t}. (40)

Using 𝒙{\bm{x}} and 𝒑{\bm{p}} (the kinetic momentum) as independent variable is more convenient. In this case, the equations of motion (33) become very simple and take the same form as Eqs. (23) and (24),

G𝒙˙\displaystyle\sqrt{G}\dot{{\bm{x}}} =\displaystyle= 𝒗p+(𝒗p𝛀)𝑩+𝑬~×𝛀,\displaystyle\bm{v}_{p}+\hbar(\bm{v}_{p}\cdot\bm{\Omega})\bm{B}+\hbar\tilde{\bm{E}}\times\bm{\Omega}, (41)
G𝒑˙\displaystyle\sqrt{G}\dot{{\bm{p}}} =\displaystyle= 𝑬~+𝒗p×𝑩+(𝑬~𝑩)𝛀.\displaystyle\tilde{\bm{E}}+\bm{v}_{p}\times\bm{B}+\hbar(\tilde{\bm{E}}\cdot\bm{B})\bm{\Omega}. (42)

where 𝑬~=𝑬+xεp\tilde{\bm{E}}=\bm{E}+\bm{\nabla}_{x}\varepsilon_{p} with εp=±|𝑲(𝒑)|+K0(𝒑)+h1(𝒙,𝒑)\varepsilon_{p}=\pm|\bm{K}({\bm{p}})|+K_{0}({\bm{p}})+h_{1}({\bm{x}},{\bm{p}}), 𝒗p=pεp{\bm{v}_{p}}=\bm{\nabla}_{p}\varepsilon_{p}, and Ωi=12ϵijkΩpjpk\Omega^{i}=\frac{1}{2}\epsilon^{ijk}\Omega_{p^{j}p^{k}}. With these equations of motion, the collisionless kinetic equation can be written as (t+𝒙˙x+𝒑˙p)f=0(\partial_{t}+\dot{{\bm{x}}}\cdot\bm{\nabla}_{x}+\dot{{\bm{p}}}\cdot\bm{\nabla}_{p})f=0 whose explicit form is same as Eq. (22):

{Gt+[𝒗p+(𝒗p𝛀)𝑩+𝑬~×𝛀]x+[𝑬~+𝒗p×𝑩+(𝑬~𝑩)𝛀]p}f(t,𝒙,𝒑)=0.\begin{split}&\Big{\{}\sqrt{G}\partial_{t}+\left[\bm{v}_{p}+\hbar(\bm{v}_{p}\cdot\bm{\Omega})\bm{B}+\hbar\tilde{\bm{E}}\times\bm{\Omega}\right]\cdot\bm{\nabla}_{x}\\ &\quad+\left[\tilde{\bm{E}}+\bm{v}_{p}\times\bm{B}+\hbar(\tilde{\bm{E}}\cdot\bm{B})\bm{\Omega}\right]\cdot\bm{\nabla}_{p}\Big{\}}f(t,{\bm{x}},{\bm{p}})=0.\end{split} (43)

Let us seek for an action in the form S=𝑑t[Ui(t,𝒙,𝒑)x˙i+Wi(t,𝒙,𝒑)p˙iξ(t,𝒙,𝒑)]S=\int dt[U^{i}(t,{\bm{x}},{\bm{p}})\dot{x}^{i}+W^{i}(t,{\bm{x}},{\bm{p}})\dot{p}^{i}-\xi(t,{\bm{x}},{\bm{p}})] for Eqs. (41) and (42). Using the least action principle, we determine Ui,WiU^{i},W^{i}, and ξ\xi as Ui=pi+Ai(t,𝒙)U^{i}=p^{i}+A^{i}(t,{\bm{x}}), Wi=ai(𝒑)W^{i}=-\hbar a^{i}({\bm{p}}), and ξ=εp+A0(t,𝒙)\xi=\varepsilon_{p}+A_{0}(t,{\bm{x}}):

S=𝑑t[(𝒑+𝑨(t,𝒙))𝒙˙𝒂(𝒑)𝒑εpA0(t,𝒙)],\displaystyle S=\int dt\left[({\bm{p}}+\bm{A}(t,{\bm{x}}))\dot{{\bm{x}}}-\hbar\bm{a}({\bm{p}})\cdot{{\bm{p}}}-\varepsilon_{p}-A_{0}(t,{\bm{x}})\right],

where 𝒂(𝒑)\bm{a}({\bm{p}}) is determined by the condition p×𝒂=𝛀\bm{\nabla}_{p}\times\bm{a}=\bm{\Omega} so that 𝒂\bm{a} is the Berry connection. Using SS we obtain the number density and current as

jμ(x)\displaystyle j^{\mu}(x) =\displaystyle= d3𝒚d3𝒑(2π)3GδS(𝒚,𝒑)δAμ(t,𝒙)f(t,𝒙,𝒑).\displaystyle-\int\frac{d^{3}\bm{y}d^{3}{\bm{p}}}{(2\pi)^{3}}\sqrt{G}\frac{\delta S(\bm{y},{\bm{p}})}{\delta A_{\mu}(t,{\bm{x}})}f(t,{\bm{x}},{\bm{p}}). (45)

The results are

j0(x)\displaystyle j^{0}(x) =\displaystyle= d3𝒑(2π)3Gf(t,𝒙,𝒑),\displaystyle\int\frac{d^{3}{\bm{p}}}{(2\pi)^{3}}\sqrt{G}f(t,{\bm{x}},{\bm{p}}), (46)
𝒋(x)\displaystyle\bm{j}(x) =\displaystyle= d3𝒑(2π)3G[𝒙˙|𝑲(𝒑)|𝛀×x]f(t,𝒙,𝒑),\displaystyle\int\frac{d^{3}{\bm{p}}}{(2\pi)^{3}}\sqrt{G}\left[\dot{{\bm{x}}}\mp\hbar|\bm{K}({\bm{p}})|\bm{\Omega}\times\bm{\nabla}_{x}\right]f(t,{\bm{x}},{\bm{p}}),

where the second term in 𝒋\bm{j} is because h1(𝒙,𝒑)h_{1}({\bm{x}},{\bm{p}}) depends on 𝑨\bm{A}. Note the similarity with equations (28) and (II). The divergence of jμj^{\mu} reads (omitting O(2)O(\hbar^{2}) terms)

μjμ\displaystyle\partial_{\mu}j^{\mu} =\displaystyle= d3𝒑(2π)3[tG+x(G𝒙˙)+p(G𝒑˙)]f\displaystyle\int\frac{d^{3}{\bm{p}}}{(2\pi)^{3}}\left[\partial_{t}\sqrt{G}+\bm{\nabla}_{x}\cdot(\sqrt{G}\dot{{\bm{x}}})+\bm{\nabla}_{p}\cdot(\sqrt{G}\dot{{\bm{p}}})\right]f (48)
=\displaystyle= 𝑬𝑩d3𝒑(2π)3p𝛀(𝒑)f(t,𝒙,𝒑),\displaystyle\hbar\bm{E}\cdot\bm{B}\int\frac{d^{3}{\bm{p}}}{(2\pi)^{3}}\bm{\nabla}_{p}\cdot\bm{\Omega}({\bm{p}})f(t,{\bm{x}},{\bm{p}}),

where we have used 333In terms of the generalized Berry curvature Ωαβ\Omega_{\alpha\beta}, it can be expressed by Gt+Gx˙ixi+Gp˙ipi=Θpixit+Θpjxjxipiεp+Θxjpjpixiεp\frac{\partial\sqrt{G}}{\partial t}+\frac{\partial\sqrt{G}\dot{x}^{i}}{\partial x^{i}}+\frac{\partial\sqrt{G}\dot{p}^{i}}{\partial p^{i}}=\Theta_{p^{i}x^{i}t}+\Theta_{p^{j}x^{j}x^{i}}\partial_{p^{i}}\varepsilon_{p}+\Theta_{x^{j}p^{j}p^{i}}\partial_{x^{i}}\varepsilon_{p} where Θαβγ=αΩβγ+βΩγα+γΩβα\Theta_{\alpha\beta\gamma}=\partial_{\alpha}\Omega_{\beta\gamma}+\partial_{\beta}\Omega_{\gamma\alpha}+\partial_{\gamma}\Omega_{\beta\alpha} is the Berry monopole charge function which is the exteriror derivative of Berry curvature Gao et al. (2021).

tG+x(G𝒙˙)+p(G𝒑˙)\displaystyle\partial_{t}\sqrt{G}+\bm{\nabla}_{x}\cdot(\sqrt{G}\dot{{\bm{x}}})+\bm{\nabla}_{p}\cdot(\sqrt{G}\dot{{\bm{p}}}) =\displaystyle= p𝛀(𝒑)𝑬~𝑩,\displaystyle\hbar\bm{\nabla}_{p}\cdot\bm{\Omega}({\bm{p}})\tilde{\bm{E}}\cdot\bm{B},

which is a direct consequence of Eqs. (41) and (42). The quantity p𝛀\bm{\nabla}_{p}\cdot\bm{\Omega} vanishes except for points where 𝛀\bm{\Omega} is singular. In fact, substituting the expression for 𝛀\bm{\Omega}, we have

p𝛀(𝒑)\displaystyle\bm{\nabla}_{p}\cdot\bm{\Omega}({\bm{p}}) =\displaystyle= det(Kapi)K𝛀K\displaystyle{\rm det}\left(\frac{\partial K^{a}}{\partial p^{i}}\right)\bm{\nabla}_{K}\cdot\bm{\Omega}_{K} (50)
=\displaystyle= ±2πdet(Kapi)δ(3)(𝑲),\displaystyle\pm 2\pi\,{\rm det}\left(\frac{\partial K^{a}}{\partial p^{i}}\right)\delta^{(3)}(\bm{K}),

where the repeated indices all summed over 131-3 and 𝛀K=±𝑲^/(2𝑲2)\bm{\Omega}_{K}=\pm\hat{\bm{K}}/(2\bm{K}^{2}) is the Berry curvature in 𝑲\bm{K}-space. To derive this relation, we have used the Schouten identity ϵijkK,j[aK,kbK,idΩKc]=0\epsilon^{ijk}K^{[a}_{,j}K^{b}_{,k}K^{d}_{,i}\Omega_{K}^{c]}=0 with K,ja=Ka/pjK^{a}_{,j}=\partial K^{a}/\partial p^{j}. The determinant in the right-hand side of Eq. (50) represents the Berry monopole charge in 𝑲\bm{K}-space and is mathematically a Jacobian for the map from 𝒑\bm{p}-space to 𝑲\bm{K}-space which makes the integral of p𝛀\bm{\nabla}_{p}\cdot\bm{\Omega} have clear topological meaning:

12πd3𝒑p𝛀(𝒑)\displaystyle\frac{1}{2\pi}\int d^{3}{\bm{p}}\bm{\nabla}_{p}\cdot\bm{\Omega}({\bm{p}}) =\displaystyle= ±𝒑N𝒑,\displaystyle\pm\sum_{\bm{p}^{*}}N_{{\bm{p}}^{*}}, (51)

where 𝒑{\bm{p}}^{*} is the points in 𝒑{\bm{p}}-space where 𝑲(𝒑)=0\bm{K}({\bm{p}}^{*})=0 (i.e., the location of Berry monopole in 𝒑{\bm{p}}-space) and N𝒑N_{{\bm{p}}^{*}} is the winding number of map 𝒑𝑲=𝑲(𝒑){\bm{p}}\rightarrow\bm{K}=\bm{K}({\bm{p}}) around 𝒑{\bm{p}}^{*}. Collecting the above results, we obtain the following relation expressing the chiral anomaly:

μjμ\displaystyle\partial_{\mu}j^{\mu} =\displaystyle= ±4π2𝑬𝑩𝒑N𝒑f(t,𝒙,𝒑).\displaystyle\pm\frac{\hbar}{4\pi^{2}}\bm{E}\cdot\bm{B}\sum_{\bm{p}^{*}}N_{{\bm{p}}^{*}}f(t,{\bm{x}},{\bm{p}}^{*}). (52)

This extends Eq. (30) to a more general case described by Hamiltonian (1). It shows that the term K0K_{0} in Hamiltonian (1) does not change the form of the anomaly relation and contributes to the chiral anomaly only through the distribution function ff. We emphasize that the relation (52) can also be understood through the index theorem Yee and Yi (2020).

To end this section, we give the expressions for CME of fermions of type (1) for an equilibrium distribution function which depends only on the energy εp\varepsilon_{p}, f=f(εp)f=f(\varepsilon_{p}),

𝒋CME\displaystyle\bm{j}_{\rm CME} =\displaystyle= 4π2𝑩𝒑N𝒑F(ε𝒑),\displaystyle\mp\frac{\hbar}{4\pi^{2}}\bm{B}\sum_{\bm{p}^{*}}N_{{\bm{p}}^{*}}F(\varepsilon_{{\bm{p}}^{*}}), (53)

where F(εp)F({\varepsilon_{p}}) satiesfies F(εp)=f(εp)F^{\prime}(\varepsilon_{p})=f(\varepsilon_{p}) and the boundary condition F(εp)0F(\varepsilon_{p})\rightarrow 0 for 𝒑{\bm{p}}\rightarrow\infty.

IV Summary and outlook

In summary, we have thoroughly examined chiral anomaly for fermions described by Hamiltonians (1) and (2) via the kinetic approach. For Hamiltonian (2), we use the Wigner function method and derive the corresponding chiral kinetic equation up to O()O(\hbar) order. Comparing to the relativistic case, the only change is the on-shell condition which constrains also the single-particle equations of motion and the distribution function. The chiral anomaly is seen to be sourced by the Berry monopole at 𝒑=𝟎{\bm{p}}=\bm{0} and is formally unchanged comparing to the relativistic case. We then extend the analysis to the more general Hamiltonian (1) by directly applying the semi-classical expansion to it. This amounts to map the O()O(\hbar) order behavior of the quantum dynamics of a spinful particle to the classical dynamics of a spinless particle equipped with a modified \hbar-dependent Hamilton function and sympletic form. The Berry monopole charge is given by the winding number of the map from 𝒑{\bm{p}}-space to 𝑲\bm{K}-space around each Berry monopole. Due to this, the chiral anomaly relation is correspondly changed [see Eq. (52)].

From the above analysis, the source of chiral anomaly from the point of view of kinetic theory is clearly displayed, showing that the chiral anomaly has an infrared origin stemming from the singularities in phase space (the Berry monopole). This could be helpful for understanding the related topological transport phenomena in non-relativistic systems. It would be also interesting to make a field-theoretical analysis for Hamiltonians (1) and (2) to reveal the connection of chiral anomaly with ultraviolet divergence in field theory. Such an analysis has been performed for Hamiltonian (1) without K0K_{0} term Yee and Yi (2020) and will be presented for the case with K0K_{0} in a future work.

Acknowledgements.
The authors thank Sahal Kaushik, Dmitri Kharzeev, Yu-Chen Liu, Satoshi Nawata, Evan Philip, Xin-Li Sheng, and Yong-Shi Wu for helpful discussions. This work is supported by NSFC under Grant No. 12075061 and Shanghai NSF under Grant No. 20ZR1404100.

References