Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2022 Mar 14;119(12):e2200367119. doi: 10.1073/pnas.2200367119

Skin effect as a probe of transport regimes in Weyl semimetals

Paweł Matus a,b, Renato M A Dantas a,b,c, Roderich Moessner a,b, Piotr Surówka a,b,d,1
PMCID: PMC8944275  PMID: 35286187

Significance

Weyl semimetals are a class of three-dimensional materials, whose low-energy excitations mimic massless fermions. In consequence they exhibit various unusual transport properties related to the presence of chiral anomalies, a subtle quantum phenomenon that denotes the breaking of the classical chiral symmetry by quantum fluctuations. In this work we present a universal description of transport in weakly disordered Weyl semimetals with several scattering mechanisms taken into account. Our work predicts the existence of a new anomaly-induced transport regime in these materials and gives a crisp experimental signature of a chiral anomaly in optical conductivity measurements. Finally, by also capturing the hydrodynamic regime of quasiparticles, our construction bridges the gap between developments in electronic fluid mechanics and three-dimensional semimetals.

Keywords: skin effect, Weyl semimetals, quantum anomalies

Abstract

We study the propagation of an oscillatory electromagnetic field inside a Weyl semimetal. In conventional conductors, the motion of the charge carriers in the skin layer near the surface can be diffusive, ballistic, or hydrodynamic. We show that the presence of chiral anomalies, intrinsic to the massless Weyl particles, leads to a hitherto neglected nonlocal regime that can separate the normal and viscous skin effects. We propose to use this regime as a diagnostic of the presence of chiral anomalies in optical conductivity measurements. These results are obtained from a generalized kinetic theory that includes various relaxation mechanisms, allowing us to investigate different transport regimes of Weyl semimetals.


Electronic transport phenomena often present particularly direct and accurate probes of the properties of conducting materials. Conventional metals are well described by the Drude theory that captures the diffusive movement of charge carriers. In very clean materials, where the scattering of electrons with impurities is not frequent, transport departs from being diffusive. Depending on the interaction strength, transport exhibits either ballistic or hydrodynamic behavior (1, 2). So far the experimental candidates to investigate such effects are two-dimensional (Al,Ga)As heterostructures (35) and graphene (612). A detailed theoretical analysis describing viscous electronics and ballistic-to-hydrodynamic cross-overs in clean two-dimensional structures is done by means of kinetic theory with various relaxation times corresponding to different scattering mechanisms (13, 14).

In a parallel line of developments, much effort has been devoted to the study of topological effects in the transport of massless quasiparticles. In three dimensions, these can be realized as low-energy electronic excitations in materials known as Dirac semimetals and Weyl semimetals. In Weyl semimetals either time-reversal symmetry or inversion symmetry is absent and the spectrum in the vicinity of the Fermi energy consists of an even number of linear band crossings known as Weyl nodes. Each crossing hosts fermions with a well-defined chirality, and the sum of the chiralities in the Brillouin zone is zero (15, 16). The fact that the Weyl nodes are generically separated in the reciprocal space makes these systems an ideal platform to investigate transport phenomena related to the chiral anomaly. It describes the breaking of the classical chiral symmetry by quantum fluctuations in parallel electric and magnetic fields (1719). At weak coupling, the imprint of the chiral anomaly has been connected to negative magnetoresistance observed in Dirac and Weyl semimetals (2028). In the hydrodynamic regime, macroscopic effects tied to chiral anomalies are the chiral magnetic effect (CME), the chiral vortical effect, and thermal transport phenomena related to gravitational anomalies (2938). Despite the above predictions, on the one hand, the direct identification of the chiral anomalies with negative magnetoresistance is problematic due to other interfering effects (39); on the other hand, the hydrodynamic regime is more difficult to control experimentally (for recent progress see refs. 40 and 41) and the corresponding effects are difficult to measure in practice.

The relevant degrees of freedom to capture transport in semimetals with a finite Fermi energy and at low temperatures can be modeled as long-lived quasiparticles. In this case the pertinent quantity is the single-particle distribution function due to the fact that we can treat interactions perturbatively. The evolution of the single-particle distribution function is given by the Boltzmann kinetic equation, which incorporates both the single-particle dynamics and the relaxation processes arising from interactions between constituents of the system. This provides the semiclassical description of physical observables that vary slowly in space and time, establishing a statistical description of a many-body electron system. The effect of collisions of the particles has to be encapsulated by an appropriate choice of the collision term, which is usually done by a phenomenological treatment of the appropriate relaxation mechanisms (42, 43). In the case of massless fermions under the influence of an external magnetic field, the classical evolution gets modified due to the Berry phase contribution. Thus, the dynamics of the distribution function also change and are given by the chiral kinetic equation (4448).

Transport properties of Weyl semimetals are not limited to steady flow configurations. In fact, semimetals driven by alternating current (ac) electric fields exhibit a rich phenomenology and reveal new effects absent in direct current (dc) settings in both two (4957) and three dimensions (5867).

Here, our goal is to investigate the ramifications of chiral anomalies on transport properties at finite frequency. One of the classic examples of a finite driving transport phenomenon is the skin effect (68). It states that an alternating electric current is distributed mainly close to the wall of the conductor and decays exponentially within the conductor. Therefore, the skin effect increases its effective resistance through the reduction of the effective cross-section of the conductor. This reduction is parameterized by a distance, called the penetration depth or the skin depth, that depends on the driving frequency. Depending on the dominant relaxation mechanism in the skin layer, which in general changes with the skin depth, three types of skin effect are conventionally distinguished: normal, anomalous, and viscous (2, 69). The normal skin effect assumes that the constitutive relation between the current and the electric field is local. The anomalous skin effect appears when this assumption is not valid and the local current depends nonlocally on the field distribution. Note that the name anomalous skin effect should not be confused with the influence of chiral anomalies. Both the normal and anomalous skin effects appear when the interactions between electrons are weak. Third, the viscous skin effect manifests itself in clean materials when the interactions between charge carriers become large and the charge flow is described by hydrodynamics. It is an example of boundary layer phenomena in fluid mechanics that arise in the immediate vicinity of a bounding surface where the effects of viscosity are significant.

The goal of this paper is to study skin effects for a Weyl metal, focusing on imprints of chiral anomalies. Chiral anomaly leads to the appearance of the chiral magnetic current (23)

JCMEτ5(E·B)B, [1]

where τ5 is the time needed for the relaxation of the axial charge (i.e., the difference between the charge densities of the particles with opposite chirality). In the standard treatment, which focuses on the dc conductivity, τ5 is related to internode scattering induced by impurities (7072). However, when an ac field is applied, different relaxation mechanisms may become dominant.

In this work we provide a detailed study of the change of the chiral magnetic conductivity across a wide frequency spectrum. To this end, we generalize the chiral kinetic theory framework to account for different relaxation processes present in realistic materials. These processes include momentum-relaxing collisions coming from the scattering of quasiparticles with impurities as well as momentum-conserving collisions between quasiparticles. We follow the path of expanding the collision kernel in the eigenfunctions of the angular momentum. Such a procedure was successfully applied in the context of two-dimensional materials (7377).

In the process we uncover a transport regime, the existence of which depends on the presence of the chiral anomaly. In this regime the relaxation of momentum is dominated by scattering off impurities, but the relaxation of the axial charge in the skin layer takes place mainly through the diffusion of particles out of the layer and into the bulk of the material. This process causes a significant increase of the skin depth, as well as a different scaling of the surface impedance with frequency, in the presence of magnetic field. This observation gives a clear experimental imprint of the chiral anomalies in optical conductivity measurements.

Transport Regimes and the Skin Effect

The skin effect is concerned with the flow of charge in a conductor under the influence of an oscillating electric field. It was first discovered in the context of spherical conductors (78) as a consequence of Maxwell’s equations, which impose the propagation of electric field in a medium to be described by the equation

2E(·E)c2t2E=μtJ, [2]

where J is the current density, c is the speed of light, and μ is the magnetic permeability in the medium. When propagating inside a conductor, the electric field is attenuated by the induced currents and the magnitude of the field drops exponentially with the distance. This decay is characterized by a length δ called the skin depth (Fig. 1). To study the skin effect, one needs to supplement Maxwell’s equations with a model describing the dynamics of particles in an electric field. The simplest model providing such dynamics is of the constitutive type and assumes a local relation between the current J(x,t) and the electric field E(x,t). This assumption is not always justified and nonlocal effects have to be taken into account, which can be done for almost free electrons by resorting to kinetic theory (79). Finally, in clean materials the relation between the current and the electric field can be hydrodynamic and controlled by the fluid viscosity. Kinetic theory can also be used to capture the effects of viscosity, providing a unified interpretation of various regimes in terms of different scattering mechanisms in the conductor, which affects the final form of the optical conductivity (80).

Fig. 1.

Fig. 1.

A schematic illustration of the skin effect. The propagation of an electromagnetic wave entering a conductor is limited to a layer of width δ, called the skin depth. The skin depth changes with the frequency of the electric field ω, making it possible to probe different transport regimes in a single sample of a Weyl semimetal. To determine the effect of the chiral anomaly on transport, the sample is placed in an external magnetic field B aligned parallel to the propagating electric field.

Electrons moving in the bulk of a (semi)metal experience different types of scattering: 1) electron–electron collisions, which conserve the total momentum, and 2) electron–impurity and electron–phonon collisions, which dissipate momentum. Correspondingly, we denote the mean free path between momentum-conserving collisions as lmc and the mean free path for momentum-relaxing collisions as lmr. Other length scales essential to the problem are the skin depth δ, as well as the path traversed by an electron over one period of oscillation of the field, lωvF/ω, where vF is the Fermi velocity and ω is the frequency of the driving electric field. Depending on the relative values of these length scales, one can distinguish the following different transport regimes:

  • 1)

    When lmrδ or lωδ, the electric field can be assumed to be uniform on the length of the free path of an electron, and consequently the relation between the electric field and electric current is local in space: J(x,ω)=σ(ω)E(x,ω). This is the regime where the usual Drude formula for optical conductivity can be used. The skin effect in this regime is called the normal skin effect (78). In this article we further differentiate between the low-frequency normal regime (when lmr is the shortest length scale) and the high-frequency normal regime (when lω is the shortest length scale).

  • 2)

    In the case of very pure samples at low temperatures, the electron–impurity and electron–phonon collisions can become less frequent than the electron–electron collisions; that is, lmclmr. Then, in some frequency range it may happen that lmcδlmr. In that case, the motion of the electron can be seen as a random walk with step size equal to lmc. The length of the path that an electron has to traverse to cross the skin layer is of the order δ2/lmc, so momentum relaxation in the skin layer happens on the length scale of min{δ2/lmc,lmr}. If the condition δ2/lmclmr is fulfilled, we are in the hydrodynamic regime, where the skin effect is known as the viscous skin effect, first studied by Gurzhi (81).

  • 3)

    Finally, when in some frequency range δ is the smallest length scale, i.e., δmin{lmr,lmc,lω}, we enter a regime in which electrons can be assumed not to undergo any scattering in the skin layer, so the motion of the electrons is ballistic. The skin effect in this regime is called the anomalous skin effect (79).

In Weyl metals the presence of the chiral anomaly complicates this picture. When the material is placed in parallel electric and magnetic fields, the axial anomaly causes an imbalance between the densities of quasiparticles with different chiralities, and the chiral magnetic effect gives rise to the current in Eq. 1. Therefore, in addition to the rate of relaxation of momentum, also the rate of relaxation of the axial charge becomes relevant for the transport of electric charge.

While at low frequencies both momentum and axial charge relaxation originate in the same physical processes, namely collisions with impurities, the relevant length scale for the latter, which we call linter, can be much longer than lmr (7072). Consequently, in a certain frequency range lmrδlinter can occur. In analogy to our discussion of the hydrodynamic regime, in addition to the internode scattering, relaxation of the axial charge takes place also through diffusion. Since lmrδ, we are in the diffusive regime and the time needed for the charge imbalance to diffuse through the skin layer is of the order δ2/(vFlmr). If the condition δ2/lmrlinter is satisfied, diffusion becomes the primary mechanism of the axial charge relaxation in place of the internode scattering. The characteristic feature of this regime is that while the classical contribution to the conductivity can still be described by the local Drude formula, the contribution resulting from the quantum anomaly is nonlocal. Therefore, we call this regime the anomaly-induced nonlocal (AIN) regime. The interplay between the classical and quantum contributions has nontrivial consequences for the skin effect.

Kinetic Theory with Second-Order Corrections

Our aim is to calculate currents in a Weyl semimetal by solving the Boltzmann kinetic equation. Hereafter, we consider a Weyl semimetal composed of two Weyl nodes with opposite chirality and assume that there are no other bands contributing to the conductivity. Consequently, we consider two distribution functions f(s) labeled by the chirality of the node s=±1. In the vicinity of the node with chirality s, and in the absence of external fields, we take the system to be described by the idealized low-energy Hamiltonian,

H(p)=svσ·p. [3]

For such a Hamiltonian, the dispersion relation for the conduction bands takes the simple form

ϵ0(p)=v|p|. [4]

Finally, our semiclassical approach assumes a system whose Fermi energy ϵF satisfies the inequalities kTϵF and ωϵF, such that there are no antiparticle excitations.

The semiclassical equations of motion in the chiral kinetic theory are

D(s)x.(s)=[vM(s)+eE×Ω(s)+e(vM(s)·Ω(s))B], [5]
D(s)p.(s)=[eE+evM(s)×B+e2(E·B)Ω(s)], [6]

where vM(s)=pϵM(s) is the group velocity and

D(s)=1+eB·Ω(s) [7]

is the volume element of the phase space in the presence of the Berry curvature Ω(s) (46, 82). Charge and current densities are defined as

ρ(x,t)=es=±1d3p(2π)3D(s)f(s)(x,p,t), [8]
J(x,t)=es=±1d3p(2π)3D(s)x.(s)f(s)(x,p,t). [9]

The distribution functions f(s)(x,p,t) can be determined from the corresponding Boltzmann equations

tf(s)+x.(s)·xf(s)+p.(s)·pf(s)=C(s)[f(s),f(s)], [10]

where C(s) is the collision integral (discussed in the next section).

The presence of external fields changes the simple dispersion relation given in Eq. 4. In particular, corrections linear in B, related to the orbital magnetic moment (OMM) of the wavepackets, have been shown to lead to a qualitative change in the magnetoconductance (83, 84). As the chiral magnetic effect is seen on the order B2, corrections to the semiclassical equations of motion quadratic in B can also potentially affect the result. For that reason, in this work we include second-order corrections to the energy (85) (SI Appendix):

ϵM(s)(p)=v[|p|seB·p2|p|2+e22|B|28|p|3e22(B·p)28|p|5]. [11]

The magnetic field also modifies the Berry curvature (85, 86) (SI Appendix), which up to linear order in reads

Ω(s)(p)=sp2|p|3eB4|p|4+e(B·p)p2|p|6. [12]

Using these formulas, the equations of motion are valid up to order B2. Second-order corrections to the equations of motion in the context of chiral kinetic theory were considered before in refs. 8789.

It is important to state that the semiclassical approach is valid only for sufficiently weak magnetic fields. To be more precise, it is assumed that the dimensionless parameter

α=e|B|v22ϵF2 [13]

satisfies α1; otherwise the analysis would have to take into account the quantization of the Landau levels. Thus, α quantifies the magnitude of quantum corrections to the classical equations of motion.

Linearized Boltzmann Equation

The collision term C(s)[f(s),f(s)] in Eq. 10 is zero when both particle species follow the Fermi–Dirac distribution; i.e., C(s)[f0(s),f0(s)]=0 with

f0(s)(p)=[exp(ϵM(s)(p)ϵFkT)+1]1. [14]

To solve the Boltzmann equation, f(s) is expanded around f0(s). Setting f(s)=f0(s)+f1(s) in Eq. 10 shows that the leading-order solution f0 contributes a term proportional to pf0(s)=pϵM(s)·ϵMf0(s) to the equation. Consequently, we parameterize the full solution as follows:

f(s)=f0(s)(p)+ϵMf0(s)·η(s)(x,p,t). [15]

For kTϵF,ϵMf0(s)δ[ϵM(s)(p)ϵF], which means that we have only to pay attention to η(s)(x,p,t) with momenta p close to the Fermi surface. The shape of the Fermi surface is determined by equating the right-hand side (RHS) of Eq. 11 to ϵF and solving for |p|. To that end, we introduce the unit vectors pointing in the direction of magnetic field and momentum,

b^=B/|B|,          p^=p/|p|, [16]

respectively. The Fermi momentum in the direction p^ can then be written, up to the second order in B, as

pF(s)(p^)=ϵFv[1+sα(p^·b^)12α212α2(p^·b^)2]. [17]

Using this result it is possible to evaluate the Berry curvature Ω(s) and the group velocity vM(s) at the Fermi surface up to the second order in α:

Ω(s)(pF(s)(p^),p^)=1e|B|[sαp^α2b^], [18]
vM(s)(pF(s)(p^),p^)=v[p^sαb^+2sα(p^·b^)p^32α2p^+α2(p^·b^)b^32α2(p^·b^)2p^]. [19]

In what follows we make the dependence on pF implicit; i.e., Ω(s)(pF(s)(p^),p^)Ω(s)(p^),vM(s)(pF(s)(p^),p^)vM(s)(p^),η(s)(pF(s)(p^),p^)η(s)(p^).

Next, we turn our attention to the collision term on the RHS of Eq. 10. We linearize this term by introducing the collision operator C^(s) defined as

C^(s)[η(s),η(s)](ϵMf0(s))1C(s)[f(s),f(s)], [20]

where f(s)=f0(s)+ϵMf0(s)·η(s). At first, we neglect possible interactions between the nodes and consider properties of the collision operator acting on a well-defined chiral state C^(s)[η(s)]C^(s)[η(s),0]. This collision operator can be shown to be Hermitian and nonpositive with respect to the inner product (73, 90, 91)

η|ζ(s)=d3p(2π)3D(s)(p)(ϵMf0(s))η(p)*ζ(p). [21]

In the absence of an external magnetic field, B=0, the Fermi surface is a sphere and D(s)(p)=1. Then, if we assume that the scattering in the bulk is isotropic, the problem is spherically symmetric and the eigenfunctions of C^(s) are the spherical harmonics

C^(s)[Ylm(p^)]=Γl,mYlm(p^)    (B=0). [22]

The eigenvalues Γl,m describe the relaxation of the different modes. For example, if the distribution function η(s) is expanded in spherical harmonics as

η(s)=l,mXlm(s)Ylm(p^), [23]

where Xlm(s)=Xlm(s)(x,t) are some functions of position and time, the rate of change of the charge density is

dρ(s)dt=ed3p(2π)3df(s)(x,p,t)dt=ed3p(2π)3ϵMf0(s)·C^(s)[η(s)]=eϵF2(2π)2π3v3Γ0,0(X00(s)). [24]

A simple calculation shows that ρ(s)X00(s). Thus, Γ0,0 describes the rate of the particle number relaxation. Similarly, using the fact that p^zY10(p^),p^x(Y11(p^)Y11(p^)), and p^yi(Y11(p^)+Y11(p^)), it is easy to see that Γ1,M, where M=1,0,1 describe the rates of relaxation of momentum.

In the case of nonzero magnetic field, we can no longer use the spherical symmetry to find the eigenfunctions of the collision operator. Nevertheless, we can assume that they should be of the form Klm(s)=Ylm+O(α) and that the corresponding eigenvalues, Γ0,0 and Γ1,M, should again describe the rates of change of ρ and p, respectively. The details of finding the functions Klm(s) can be found in SI Appendix. In this work we expand Klm(s) in spherical harmonics to the second order in α.

Last, let us construct a collision operator that includes the relevant relaxation processes. One of them is the internode scattering, which transfers particles between the two nodes, leading to a dissipation of the axial charge; we denote the frequency at which this process happens by Γinter. Another process is the intranode scattering that, together with the internode scattering, relaxes momentum; we denote the total rate at which momentum is dissipated by Γmr and, consequently, we take Γ1,M=Γmr for M=1,0,+1. Finally, we include in our analysis electron–electron scattering, which conserves momentum. Because modes with a high angular momentum do not contribute to the current, and to simplify the calculations, we assume a constant relaxation rate Γtot for all modes with the total angular momentum L2. All the scattering mechanisms that we consider, i.e., the inter- and intranode scattering, as well as the electron–electron collisions, contribute to Γtot. On the basis of the physical origin of the relaxation rates we take Γinter<Γmr<Γtot. Introducing the operators

P0(s)=|K00(s)K00(s)||K00(s)K00(s)|,P1(s)=M=1,0,1|K1M(s)K1M(s)|,Phigher(s)=1|K00(s)K00(s)|P1(s), [25]

where s defines the distribution function that the operator acts on, the collision operator can be written as

C^(s)=ΓinterP0(s)ΓmrP1(s)ΓtotPhigher(s). [26]

Similar expansions of the collision operator were used in refs. 7477 for different systems. It can be verified that with this choice of collision operator ddtρ(s)Γinter(ρ(s)ρ(s)) and, in the absence of external force and internode scattering, ddtp(s)Γmrp(s).

We are now in a position to write down the linearized Boltzmann equation (Eq. 10). Recalling that we need to compute η(s) up to linear order in E and quadratic in B, we can neglect the anomalous Hall term in the spatial-derivative term

D(s)(p^)x.(s)·xf(s)=ϵMf0(s)[vM(s)+e(vM(s)·Ω(s))B]·xη(s), [27]

while the relevant contributions of the momentum-derivative term are

D(s)(p^)p.(s)·pf(s)=[eE+e2(E·B)Ω(s)]·vM(s)ϵMf0+e[vM(s)×B]·pη(s)ϵMf0(s). [28]

Note that we use the formulas for Ω(s) and vM(s) given in Eqs. 18 and 19, respectively, to account for the second-order corrections to the equations of motion. The last term on the RHS of Eq. 28 results from the Lorentz force and is the source of the classical Hall effect and classical magnetoresistance. Because these phenomena are well known and this article focuses on quantum corrections to the longitudinal conductivity, we do not include them here. Additionally, we are going to consider a setup where the electric field is parallel to the (external) magnetic field, in which case the classical magnetoresistance vanishes in the normal regime. On the other hand, in the hydrodynamic and ballistic regimes the situation is more complicated: For details, see SI Appendix. These considerations allow us to write the linearized Boltzmann equation

D(s)tη(s)+[vM(s)+e(vM(s)·Ω(s))B]·xη(s)+e[E·vM(s)+e(E·B)(Ω(s)·vM(s))]=D(s)C^(s)[η(s),η(s)], [29]

with C^(s) defined in Eq. 26.

Bulk Conductivities

In this section, we calculate the current flowing through the bulk of a Weyl semimetal subject to a weak ac electric field. The material is modeled as an infinite half-space z  >  0, while z  <  0 is the vacuum. We place the material in a static external magnetic field parallel to the surface, B=Bx^, and consider an oscillating electric field propagating through the bulk and polarized in the same direction as the magnetic field: E(z,t)=E(z,t)x^. The electric field is assumed to be uniform in the xy plane.

The general expression for the current, after taking into account the orbital magnetization, can be written as

J=Jkin+Jmagn, [30]

where Jkin is given by Eq. 9 and represents the “kinematic” part resulting from the forward motion of the wavepackets, while

Jmagn=×s=±1d3p(2π)3D(s)(p)sev2|p|p^f(s)(q,p) [31]

is the magnetization current. However, in the setup that we consider Jmagn=0 (see SI Appendix for a proof). Consequently, in what follows

J=Jkin. [32]

Eq. 9 is now expanded to the second order in α. In this article we neglect the anomalous Hall effect, i.e., the eE×Ω(s) term. In a time-reversal symmetry-breaking Weyl semimetal with a single pair of nodes this can be justified, if the electric field is aligned parallel (or near parallel) to the vector connecting the Weyl nodes in reciprocal space. After switching to spherical coordinates (p,θ,ϕ) in momentum space, one obtains

J(z)=es=±1d3p(2π)3D(p)x.(p)(δ[ϵM(p)ϵF])η(z,p)=eϵF2v2(2π)3s=±1d(cosθ)dϕ[p^+3sα(p^·b^)p^α2p^+α2(p^·b^)b^]η(s)(z,p^), [33]

where we suppress the (s) labels on all quantities in the first line.

To find the current from Eq. 33, the distribution functions η(s)(z,p) have to be determined from the Boltzmann equation (Eq. 29). We shall work with Fourier-transformed quantities defined as follows:

η(s)(z,p,t)=dωdqeiωtiqzη(s)(q,p,ω),E(z,p,t)=dωdkeiωtiqzE(q,p,ω). [34]

The Boltzmann equation is solved perturbatively in α after expanding the out-of-equilibrium distribution in spherical harmonics as

η(s)=l,m(Alm+sαBlm+α2Clm)Ylm(p^). [35]

Next, the collision integral on the RHS of Eq. 29 is evaluated using the definition of the inner product 21; the expansion of the collision operator into projection operators 25, 26; and the form of the eigenvectors Klm(s) determined in SI Appendix. Then, Eq. 29 is projected onto a spherical harmonic of degree L and order M by multiplying both sides by YLM*(p^) and integrating over the angles. These integrals can be evaluated analytically, producing an infinite system of equations labeled by L,M. Finally, terms that are of the same order in α are equated, leading to three systems of equations (for terms of order 1, α, and α2) that are solved in turn.

We now solve the system of equations corresponding to the lowest level of approximation (α=0) to demonstrate the general strategy of solution, which is in the same spirit as the one used in ref. 92 for a two-dimensional (2D) system. At this level, the equations are*

iωA00iqv33A10=0, [36]
(iω+Γmr)A10iqv[33A00+415A20]=0, [37]
(iω+Γmr)A1±1iqv55A2±1ev2π3E=0, [38]
(iω+Γtot)ALMiqv12[AL1M+AL+1M]=0      (for L2). [39]

Note that in the above, due to the cylindrical symmetry of the system in the absence of an electric field, only equations for a fixed M are coupled. Eq. 39 is a recurrence relation that can be solved by the ansatz ALM=rL1A1M. There are two solutions, from which we choose the one that ensures convergence of the series,

r=iqv[Γtot+iω(Γtot+iω)2+q2v2], [40]

where we take the square root in the bracket to return a value with a positive real and a positive imaginary part. Inserting A2M=rA1M for M=1,0,+1 in Eqs. 3638 yields a closed system of equations that can easily be solved, while for |M|>1 the solution is trivial: ALM=0. This way, the full distribution function at the classical level is found to be

A1±1=±2π3evEΓmr+iω55iqvr,AL±1=rL1A1±1,      ALM±1=0. [41]

The same procedure can be used to find corrections to η(s) at higher orders in α as well. First, the solution for ALM is plugged into the equations that are linear in α to find BLM. Subsequently, the equations for high L are expressed as a recurrence relation that is then solved, and its solution is used to turn the system of equations for L = 0, 1, 2 into a closed system. After solving this system, the solutions for ALM and BLM are plugged into the quadratic-order equations to find CLM and the procedure is repeated. As the calculations are lengthy and not particularly instructive, we relegate the details to SI Appendix.

After finding the distributions η(s)(q,p^,ω), the current is evaluated using Eq. 33. The current can be expressed as J(q,ω)=σ(q,ω)E(q,ω), where the conductivity σ is a complicated function of q and ω. Here we present only the formulas for the conductivity in the limiting cases, in which some of the frequency scales involved in the problem are much greater than the others. The relevant frequency scales are the relaxation rates, which satisfy the inequalities Γinter<Γmr<Γtot, ω (the driving frequency), and qv. As both qv and ω can in general be placed between any two relaxation rates in this chain of inequalities, there are in principle a large number of limiting cases that can be considered. Instead of analyzing each one of them, we focus only on the regimes that could be accessed in a skin effect experiment, which we identify in the following way. We choose certain realistic material parameters, close to those reported for WP2 (40, 93): v=1.4×105 m/s, ϵF=20 meV, Γmr=109 Hz, Γtot=1010 Hz. We moreover assume that Γinter is two orders of magnitude smaller than Γmr as reported for TaAs (26), so Γinter=107 Hz. Next, let us note that the classical conductivity can be calculated using Eqs. 41 and 33 evaluated at the classical (i.e., α  =  0) level, yielding

σcl(q,ω)=2εωP21Γmr+iω55iqvr, [42]

where the plasma frequency ωP is defined by

ωP2=e2ϵF26π2ε3v [43]

and ε is the electric permittivity of the medium. Then, we can solve the Fourier-transformed Eq. 2 as

(q2+ω2c2)=iμωσ(q,ω)cl [44]

to obtain the wavectors q, which correspond to modes that can propagate through the bulk, as a function of ω. This equation has in general six solutions, but we can restrict our attention to those with a negative imaginary part and a positive real part, which in accordance with Eq. 34 correspond to modes propagating in the positive direction along z and decaying with the distance. It turns out that out of the six solutions one satisfies this condition for all ω, and one satisfies it at high frequencies only; we denote these solutions q1 and q2, respectively. The characteristic length scales at which these modes decay, defined as δi=Im[qi]1, can be used as a proxy for the classical skin depth and they are plotted in Fig. 2. Based on the different scaling of qi with ω we can identify four relevant regimes, which we call low-frequency normal, hydrodynamic, ballistic, and high-frequency normal. In all these regimes, except the last one, ω is smaller than all the relevant frequency scales and can thus be neglected in the analysis.

Fig. 2.

Fig. 2.

Dependence of the characteristic length scales at which a propagating electric field decays on ω. The two curves represent two solutions of Maxwell’s equations in the form of Eq. 44 supplemented by the formula for the classical conductivity Eq. 42 and experimentally realistic parameters (main text): v=1.4×105 m/s, ϵF=20 meV, Γmr=109 Hz, and Γtot=1010 Hz. We select only solutions satisfying Re[qi]>0,Im[qi]<0 and plot their characteristic decay lengths, defined as δi=Im[qi]1. At low frequencies there exists only one solution q1 satisfying these conditions. We can observe three transport regimes (identified as low-frequency normal, hydrodynamic, and ballistic) where δ1 shows different scaling with ω, and the cross-overs (marked with dashed black lines) happen around the values of ω for which |q1v|Γmr and |q1v|Γtot. Above a certain value of ω, around the point when |q1v|ω, another solution q2 becomes relevant. Since δ2 is the longer of the two decay lengths, it dominates the propagation of the electric field and we are in what we call the high-frequency normal regime.

When quantum effects are included, the low-frequency normal regime splits into two, as discussed in the Introduction. The different regimes and the corresponding conductivities are assembled in Table 1. It can be seen that while in both the low-frequency normal and the anomaly-induced nonlocal regimes the relaxation of momentum and axial charge are related to different frequency scales, in the other regimes they both happen at a comparable rate.

Table 1.

Conductivity and surface impedance in the different regimes and their regions of validity

Regime Valid when Conductivity σ(q,ω) Surface impedance Z(ω)
Low-frequency normal q2v2ΓinterΓmr ωΓinter εωP2[2Γmr+α23Γinterα266151Γmr] eiπ/4μωc2ωP2(2Γmr+α23Γinterα266151Γmr)12
Anomaly-induced nonlocal ΓinterΓmrq2v2Γmr2 ωq2v2/Γmr εωP2[2Γmr+α218Γmrq2v2] See Eq. 56
Hydrodynamic ΓmrΓtotq2v2Γtot2 ωq2v2/Γtot εωP2Γtotq2v2[9.026α2] 12(eiπ/8+ei5π/8)μ(ω3v2c2ωP2Γtot[9.026α2])1/4
Ballistic Γtotqv ωqv εωP21qv[4.5+4.3α2] 439eiπ/3μ(ω2vc2ωP2[4.5+4.3α2])1/3
High-frequency normal qvω Γtotω iεωP21ω[2+85α2] iμωcωP(2+85α2)12

The numerical coefficients in the hydrodynamic and ballistic regimes are rounded up to two significant numbers.

We can extract further information as to the origin of the numerical coefficients in Table 1 if before the start of the calculations we multiply the term corresponding to the OMM in Eq. 11 (that is, seB·p2|p|2) by a factor ξ1, and similarly we multiply all the terms resulting from the second-order corrections to the energy and Berry curvature in Eqs. 11 and 12 by ξ2, and then we keep track of these coefficients during the computations. This allows us to determine the following:

  • The second-order corrections to the equations of motion have no influence on the conductivity either in the normal or the AIN regimes, affecting only the hydrodynamic and ballistic results.

  • For a relatively high Γinter, namely Γinter>15Γmr/220.7Γmr, the magnetoconductivity in the low-frequency regime changes sign from positive to negative, in qualitative agreement with refs. 83 and 84. This is found to be caused by the contribution of the OMM, as setting ξ1=0 and neglecting electron–electron collisions (so that Γtot=Γmr) would result in a different formula for the conductivity, denoted here as σ˜,
    σ˜(q,ω)=εωP2[2Γmr+α23Γinterα242151Γmr], [45]
    and the magnetoconductivity would always be positive, again in agreement with refs. 83 and 84.
  • In the hydrodynamic regime the large negative coefficient is also due to the OMM as setting ξ1=0 would result in a positive coefficient
    σ˜(q,ω)εωP2Γtotq2v2[9.0+5.2α2]. [46]
  • In contrast, magnetoconductivity in the AIN regime is not affected by either the OMM or the second-order corrections and results solely from the leading-order term in the Berry curvature. The nonlocality of the quantum correction to the conductivity can be seen in its dependence on q, absent in the classical part.

Skin Effect

To find the skin depths and the surface impedance in the regimes identified in the previous section, we have to include the presence of a boundary in our analysis. This is done by imposing appropriate boundary conditions on the distribution functions. As a particularly simple choice we impose specular boundary conditions, meaning that during reflection the transverse components of velocity stay the same but the perpendicular component gets inverted; i.e., η(s)(z=0,vx,vy,vz)=η(s)(z=0,vx,vy,vz). These boundary conditions can be naturally implemented if, following ref. 42, we add a mirror-symmetric counterpart (for z  <  0) to the original (physical) half-space system. Then, in the augmented setup the condition of specular reflection means that the particles move freely across the z = 0 boundary (Fig. 3). Because the electric field is parallel to the surface, its mirror-reflected components are related by E(z,t)=E(z,t). Furthermore, the external magnetic field B does not change under the mirror reflection, which allows us to use a single formula for the conductivity in the whole space.

Fig. 3.

Fig. 3.

Implementation of specular boundary conditions. To the original system located in z  >  0 its mirror-reflected version is added in z  <  0, which creates a discontinuity in the derivative of the electric field at the boundary. A specular reflection of a particle can be seen as its mirror-reflected counterpart freely crossing the boundary.

From Maxwell’s equations we obtain the relation (using E(z,t)=E(z,t)x^)

z2E(z,t)=1c2t2E(z,t)+μtJ(z,t). [47]

Fourier transforming the left-hand side (LHS),

dz2πE(z)eiqz=dz2π(E(z)eiqz)dziqdz2πE(z)eiqzdz. [48]

The first derivative of E(z) has a discontinuity at z = 0, so that E(z=0)=E(z=0+), which means that the integration of the first term should be done separately in (,0) and (0,). The second term can be integrated by parts once again, producing

dz2πE(z)eiqz=1πE(z=+0)q2E(q). [49]

The Fourier transform of J(z,ω) is σ(q,ω)E(q,ω), where σ(q,ω) was determined in the previous section, so from Eq. 47 we obtain

E(q,ω)=E(z=0+)π1q2ω2c2+iμωσ(|q|,ω), [50]

where we explicitly write |q| because conductivity has to be an even function of q. Finally, we obtain

E(z)=E(z=0+)πdqeiqzq2ω2c2+iμωσ(|q|,ω)=2E(z=0+)π0dqeiqzq2ω2c2+iμωσ(q,ω). [51]

This result allows us to calculate surface impedance, defined as (94)

Z=E(z=0)0J(z)dz. [52]

Using Maxwell’s equations this formula can be rewritten as

Z=iμωE(z=0)E(z=0+). [53]

As the conductivities in all the regimes except AIN show only quantitative changes with respect to their classical counterparts, we restrict ourselves to presenting the impedances obtained from Eqs. 53 and 51 in Table 1.

Skin Effect in the AIN Regime.

We now present a detailed analysis of the skin effect in the most interesting case of the anomaly-induced nonlocal regime. In fact, we can also capture the crossover from the low-frequency normal regime by approximating the conductivity as

σ(q,ω)σ0[1+α2Γmr2Γinter/3+q2v2/9Γmr], [54]

where σ0=2ϵωP2/Γmr. This approximation for the conductivity works only when ΓinterΓmr,qvΓmr, and ωq2v2/Γmr. We want to evaluate the integral in Eq. 51, which can be done using standard contour integration methods. The resulting impedance in both the low-frequency normal and the AIN regimes can be seen in Fig. 4.

Fig. 4.

Fig. 4.

Dependence of the real part of the surface impedance Z on frequency ω in the low-frequency normal and the AIN regimes, obtained from Eqs. 53, 51, and 54 in the cases of no magnetic field (α  =  0) and a nonzero magnetic field (α=0.1). The dashed line is plotted on the basis of the impedance of the low-frequency normal regime as found in Table 1, and it is the result that would be measured in the absence of the nonlocal behavior. The material parameters are v=1.4×105 m/s, ϵF=20 meV, Γmr=109 Hz, and Γinter=107 Hz. The transition to the AIN regime takes place around ω=105 Hz. We see that changing the external magnetic field leads to a noticeable change of the surface impedance, especially shortly before the onset of the AIN regime, and moreover in this regime the scaling with frequency changes with respect to the no-field result.

The AIN regime can be observed when (ωμσ0)v2ΓmrΓinter. For z = 0 we integrate over the contour in the lower half-plane, and thus we are interested in the poles with a negative imaginary part. There are two such poles: q1eiπ/4μωσ0+iα29Γmr2v2 and q2iΓmr(6Γinter+9α2Γmr)/v. Please note that |q2||q1|. The electric field is

E(z)=iE(z=0+)[exp(iq1z)q1+α29Γmr2q12v2exp(iq2z)q2]. [55]

This result shows that there is an anomalous component of the electric field with the skin depth much larger than the classical skin depth, so at large distances the anomalous component dominates over the classical one. The impedance is then

Z=eiπ/4μωσ0+iα29Γmr2μωv2α29Γmrσ0v(6ΓinterΓmr+9α2)12. [56]

Therefore, in a finite magnetic field the impedance in this regime no longer scales with the square root of ω, as seen in Fig. 4. This regime with its characteristic impedance is a central finding of the present article.

Discussion and Conclusions

We provided a detailed analysis of the chiral magnetic conductivity across various transport regimes. To do so, we constructed a collision operator that captures three different relaxation mechanisms by exploiting its algebraic properties. We also included in our analysis second-order semiclassical corrections. The computed conductivities show a quantitative influence of these corrections in the hydrodynamic and anomalous regimes, although it has to be noted that due to the coexistence of classical magnetoconductivity in these regimes, the quantum corrections could be difficult to separate out experimentally.

More importantly, we uncovered an entirely different transport regime where the conductivity is a combination of a classical local part and a nonlocal quantum part. This leads to a change of scaling of the surface impedance with the driving frequency when the material is placed in an external magnetic field. This phenomenon is related to a significant increase in the skin depth, which might be possible to observe in thin enough samples.

In the end, we note that there are various other phenomena that affect transport in Weyl semimetals that we did not include in our analysis. We neglected the influence of the Fermi arcs and the anomalous Hall effect (which plays a significant role in the propagation of light in time-reversal symmetry-breaking Weyl semimetals), as well as all the effects nonlinear in the electric field. Nonetheless, the results of this article hold for time-reversal symmetry-breaking Weyl semimetals placed in weak fields when the positions of the nodes are properly aligned with the electric field, as well as (qualitatively) for noncentrosymmetric Weyl semimetals. In particular, the presence of the anomaly-induced nonlocal regime should be discernible experimentally.

Our work serves as a starting point for detailed analyses of transport regimes in Weyl semimetals, including different scattering mechanisms relevant for realistic experimental situations. As a result, one can construct a detailed theoretical description of a generation of transport measurements beyond two-dimensional analogs. This allows one to investigate effects absent in two dimensions, such as chiral vortical effect or anomalous thermal transport. It also permits one to include realistic boundary conditions, important in electron hydrodynamics and usually neglected in high-energy applications of chiral kinetic theory.

Note.

When this article was being prepared, a preprint appeared (95), in which the existence of a nonlocal regime with the same qualitative features as the AIN regime was predicted. While ref. 95, contrary to the present work, assumes ωΓinter, it is worth noting that when Γinter is replaced with iω/2 in Eq. 55, the results of the two articles are found to be in good agreement.

Supplementary Material

Supplementary File
pnas.2200367119.sapp.pdf (339.7KB, pdf)

Acknowledgments

We acknowledge discussions with Francisco Peña-Benitez, Peng Rao, and Jörg Schmalian. We thank Pavlo Sukhachov for comments on the manuscript. This work was in part supported by the Deutsche Forschungsgemeinschaft through the Leibniz Program, the cluster of excellence ct.qmat (Exzellenzcluster [EXC] 2147, Project 39085490), and Sonderforschungsbereich (SFB) 1143 (Project 247310070). P.S. was also supported by the Narodowe Centrum Nauki Sonata Bis Grant 2019/34/E/ST3/00405.

Footnotes

The authors declare no competing interest.

This article is a PNAS Direct Submission.

*The factor of 1/2 in front of the bracket in the last equation is in fact an approximate value, but this approximation works very well here. For a discussion of this approximation, see SI Appendix.

This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.2200367119/-/DCSupplemental.

Data Availability

There are no data underlying this work.

References

  • 1.Gurzhi R. N., Minimum of resistance in impurity-free conductors. Sov. Phys. JETP 44, 771–772 (1963). [Google Scholar]
  • 2.Gurzhi R. N., Hydrodynamic effects in solids at low temperature. Sov. Phys. Usp. 11, 255–270 (1968). [Google Scholar]
  • 3.Molenkamp L., de Jong M., Observation of Knudsen and Gurzhi transport regimes in a two-dimensional wire. Solid-State Electron. 37, 551–553 (1994). [Google Scholar]
  • 4.de Jong M. J., Molenkamp L. W., Hydrodynamic electron flow in high-mobility wires. Phys. Rev. B Condens. Matter 51, 13389–13402 (1995). [DOI] [PubMed] [Google Scholar]
  • 5.Gusev G. M., Levin A. D., Levinson E. V., Bakarov A. K., Viscous transport and Hall viscosity in a two-dimensional electron system. Phys. Rev. B 98, 161303 (2018). [Google Scholar]
  • 6.Crossno J., et al., Observation of the Dirac fluid and the breakdown of the Wiedemann-Franz law in graphene. Science 351, 1058–1061 (2016). [DOI] [PubMed] [Google Scholar]
  • 7.Bandurin D. A., et al., Negative local resistance caused by viscous electron backflow in graphene. Science 351, 1055–1058 (2016). [DOI] [PubMed] [Google Scholar]
  • 8.Bandurin D. A., et al., Fluidity onset in graphene. Nat. Commun. 9, 4533 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Sulpizio J. A., et al., Visualizing Poiseuille flow of hydrodynamic electrons. Nature 576, 75–79 (2019). [DOI] [PubMed] [Google Scholar]
  • 10.Berdyugin A. I., et al., Measuring Hall viscosity of graphene’s electron fluid. Science 364, 162–165 (2019). [DOI] [PubMed] [Google Scholar]
  • 11.Ku M. J. H., et al., Imaging viscous flow of the Dirac fluid in graphene. Nature 583, 537–541 (2020). [DOI] [PubMed] [Google Scholar]
  • 12.Keser A. C., et al., Geometric control of universal hydrodynamic flow in a two-dimensional electron fluid. Phys. Rev. X 11, 031030 (2021). [Google Scholar]
  • 13.Narozhny B. N., Gornyi I. V., Mirlin A. D., Schmalian J., Hydrodynamic approach to electronic transport in graphene. Ann. Phys. 529, 1700043 (2017). [Google Scholar]
  • 14.Lucas A., Fong K. C., Hydrodynamics of electrons in graphene. J. Phys. Condens. Matter 30, 053001 (2018). [DOI] [PubMed] [Google Scholar]
  • 15.Armitage N. P., Mele E. J., Vishwanath A., Weyl and Dirac semimetals in three-dimensional solids. Rev. Mod. Phys. 90, 015001 (2018). [Google Scholar]
  • 16.Nielsen H., Ninomiya M., Absence of neutrinos on a lattice: (i). Proof by homotopy theory. Nucl. Phys. B 185, 20–40 (1981). [Google Scholar]
  • 17.Nielsen H., Ninomiya M., The Adler-Bell-Jackiw anomaly and Weyl fermions in a crystal. Phys. Lett. B 130, 389–396 (1983). [Google Scholar]
  • 18.Bertlmann R. A., Anomalies in Quantum Field Theory (International Series of Monographs on Physics, Oxford University Press, Oxford, UK, 2000). [Google Scholar]
  • 19.Landsteiner K., Notes on anomaly induced transport. Acta Phys. Pol. B 47, 2617 (2016). [Google Scholar]
  • 20.Huang X., et al., Observation of the chiral-anomaly-induced negative magnetoresistance in 3D Weyl semimetal TaAs. Phys. Rev. X 5, 031023 (2015). [Google Scholar]
  • 21.Li C. Z., et al., Giant negative magnetoresistance induced by the chiral anomaly in individual Cd3As2 nanowires. Nat. Commun. 6, 10137 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Li H., et al., Negative magnetoresistance in Dirac semimetal Cd3As2. Nat. Commun. 7, 10301 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Li Q., et al., Chiral magnetic effect in ZrTe5. Nat. Phys. 12, 550–554 (2016). [Google Scholar]
  • 24.Kim H. J., et al., Dirac versus Weyl fermions in topological insulators: Adler-Bell-Jackiw anomaly in transport phenomena. Phys. Rev. Lett. 111, 246603 (2013). [DOI] [PubMed] [Google Scholar]
  • 25.Xiong J., et al., Evidence for the chiral anomaly in the Dirac semimetal Na3Bi. Science 350, 413–416 (2015). [DOI] [PubMed] [Google Scholar]
  • 26.Zhang C. L., et al., Signatures of the Adler-Bell-Jackiw chiral anomaly in a Weyl fermion semimetal. Nat. Commun. 7, 10735 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Liang S., et al., Experimental tests of the chiral anomaly magnetoresistance in the Dirac-Weyl semimetals na3bi and GdPtBi. Phys. Rev. X 8, 031002 (2018). [Google Scholar]
  • 28.Dantas R. M. A., Peña-Benitez F., Roy B., Surówka P., Magnetotransport in multi-Weyl semimetals: A kinetic theory approach. J. High Energy Phys. 2018, 69 (2018). [Google Scholar]
  • 29.Fukushima K., Kharzeev D. E., Warringa H. J., Chiral magnetic effect. Phys. Rev. D 78, 074033 (2008). [DOI] [PubMed] [Google Scholar]
  • 30.Kharzeev D. E., Warringa H. J., Chiral magnetic conductivity. Phys. Rev. D 80, 034028 (2009). [Google Scholar]
  • 31.Banerjee N., et al., Hydrodynamics from charged black branes. J. High Energy Phys. 2011, 94 (2011). [Google Scholar]
  • 32.Erdmenger J., Haack M., Kaminski M., Yarom A., Fluid dynamics of R-charged black holes. J. High Energy Phys. 2009, 055 (2009). [Google Scholar]
  • 33.Neiman Y., Oz Y., Relativistic hydrodynamics with general anomalous charges. J. High Energy Phys. 2011, 23 (2011). [Google Scholar]
  • 34.Son D. T., Surówka P., Hydrodynamics with triangle anomalies. Phys. Rev. Lett. 103, 191601 (2009). [DOI] [PubMed] [Google Scholar]
  • 35.Landsteiner K., Megías E., Peña-Benitez F., Gravitational anomaly and transport phenomena. Phys. Rev. Lett. 107, 021601 (2011). [DOI] [PubMed] [Google Scholar]
  • 36.Lucas A., Davison R. A., Sachdev S., Hydrodynamic theory of thermoelectric transport and negative magnetoresistance in Weyl semimetals. Proc. Natl. Acad. Sci. U.S.A. 113, 9463–9468 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Gorbar E. V., Miransky V. A., Shovkovy I. A., Sukhachov P. O., Hydrodynamic electron flow in a Weyl semimetal slab: Role of Chern-Simons terms. Phys. Rev. B 97, 205119 (2018). [Google Scholar]
  • 38.Dantas R. M. A., Peña-Benitez F., Roy B., Surówka P., Non-abelian anomalies in multi-Weyl semimetals. Phys. Rev. Res. 2, 013007 (2020). [Google Scholar]
  • 39.Ong N. P., Liang S., Experimental signatures of the chiral anomaly in Dirac–Weyl semimetals. Nat. Rev. Phys 3, 394–404 (2021). [Google Scholar]
  • 40.Gooth J., et al., Thermal and electrical signatures of a hydrodynamic electron fluid in tungsten diphosphide. Nat. Commun. 9, 4093 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Vool U., et al., Imaging phonon-mediated hydrodynamic flow in WTe2. Nat. Phys. 17, 1216–1220 (2021). [Google Scholar]
  • 42.Lifshitz E. M., Pitaevskii L. P., Physical Kinetics (Pergamon, 1981). [Google Scholar]
  • 43.Soto R., Kinetic Theory and Transport Phenomena (Oxford University Press, 2016). [Google Scholar]
  • 44.Loganayagam R., Surówka P., Anomaly/transport in an Ideal Weyl gas. J. High Energy Phys. 2012, 97 (2012). [Google Scholar]
  • 45.Son D. T., Yamamoto N., Berry curvature, triangle anomalies, and the chiral magnetic effect in Fermi liquids. Phys. Rev. Lett. 109, 181602 (2012). [DOI] [PubMed] [Google Scholar]
  • 46.Stephanov M. A., Yin Y., Chiral kinetic theory. Phys. Rev. Lett. 109, 162001 (2012). [DOI] [PubMed] [Google Scholar]
  • 47.Chen J. Y., Son D. T., Stephanov M. A., Yee H. U., Yin Y., Lorentz invariance in chiral kinetic theory. Phys. Rev. Lett. 113, 182302 (2014). [DOI] [PubMed] [Google Scholar]
  • 48.Chen J. Y., Son D. T., Stephanov M. A., Collisions in chiral kinetic theory. Phys. Rev. Lett. 115, 021601 (2015). [DOI] [PubMed] [Google Scholar]
  • 49.Yoshikawa N., Tamaya T., Tanaka K., High-harmonic generation in graphene enhanced by elliptically polarized light excitation. Science 356, 736–738 (2017). [DOI] [PubMed] [Google Scholar]
  • 50.Moessner R., Surówka P., Witkowski P., Pulsating flow and boundary layers in viscous electronic hydrodynamics. Phys. Rev. B 97, 161112 (2018). [Google Scholar]
  • 51.Semenyakin M., Falkovich G., Alternating currents and shear waves in viscous electronics. Phys. Rev. B 97, 085127 (2018). [Google Scholar]
  • 52.Hafez H. A., et al., Extremely efficient terahertz high-harmonic generation in graphene by hot Dirac fermions. Nature 561, 507–511 (2018). [DOI] [PubMed] [Google Scholar]
  • 53.Moessner R., Morales-Durán N., Surówka P., Witkowski P., Boundary-condition and geometry engineering in electronic hydrodynamics. Phys. Rev. B 100, 155115 (2019). [Google Scholar]
  • 54.Alekseev P. S., Magnetic resonance in a high-frequency flow of a two-dimensional viscous electron fluid. Phys. Rev. B 98, 165440 (2018). [Google Scholar]
  • 55.Chandra M., Kataria G., Sahdev D., Sundararaman R., Hydrodynamic and ballistic AC transport in two-dimensional Fermi liquids. Phys. Rev. B 99, 165409 (2019). [Google Scholar]
  • 56.Levchenko A., Schmalian J., Transport properties of strongly coupled electron–phonon liquids. Ann. Phys. 419, 168218 (2020). [Google Scholar]
  • 57.McIver J. W., et al., Light-induced anomalous Hall effect in graphene. Nat. Phys. 16, 38–41 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Sodemann I., Fu L., Quantum nonlinear Hall effect induced by berry curvature dipole in time-reversal invariant materials. Phys. Rev. Lett. 115, 216806 (2015). [DOI] [PubMed] [Google Scholar]
  • 59.Morimoto T., Zhong S., Orenstein J., Moore J. E., Semiclassical theory of nonlinear magneto-optical responses with applications to topological Dirac/Weyl semimetals. Phys. Rev. B 94, 245121 (2016). [Google Scholar]
  • 60.Burkov A. A., Dynamical density response and optical conductivity in topological metals. Phys. Rev. B 98, 165123 (2018). [Google Scholar]
  • 61.Osterhoudt G. B., et al., Colossal mid-infrared bulk photovoltaic effect in a type-I Weyl semimetal. Nat. Mater. 18, 471–475 (2019). [DOI] [PubMed] [Google Scholar]
  • 62.Golub L. E., Ivchenko E. L., Spivak B., Semiclassical theory of the circular photogalvanic effect in gyrotropic systems. Phys. Rev. B 102, 085202 (2020). [Google Scholar]
  • 63.Kovalev S., et al., Non-perturbative terahertz high-harmonic generation in the three-dimensional Dirac semimetal Cd3As2. Nat. Commun. 11, 2451 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Cheng B., et al., Efficient terahertz harmonic generation with coherent acceleration of electrons in the dirac semimetal Cd3As2. Phys. Rev. Lett. 124, 117–402 (2020). [DOI] [PubMed] [Google Scholar]
  • 65.Sukhachov P. O., Gorbar E. V., Stray magnetic field and stability of time-dependent viscous electron flow. Phys. Rev. B 104, 195111 (2021). [Google Scholar]
  • 66.Weber C. P., Ultrafast investigation and control of Dirac and Weyl semimetals. J. Appl. Phys. 129, 070901 (2021). [Google Scholar]
  • 67.Dantas R. M. A., Wang Z., Surówka P., Oka T., Nonperturbative topological current in Weyl and Dirac semimetals in laser fields. Phys. Rev. B 103, L201105 (2021). [Google Scholar]
  • 68.Tanner D. B., Optical Effects in Solids (Cambridge University Press, 2019). [Google Scholar]
  • 69.Kaganov M. I., Lyubarskiy G. Y., Mitina A. G., The theory and history of the anomalous skin effect in normal metals. Phys. Rep. 288, 291–304 (1997). [Google Scholar]
  • 70.Son D. T., Spivak B. Z., Chiral anomaly and classical negative magnetoresistance of Weyl metals. Phys. Rev. B 88, 104412 (2013). [Google Scholar]
  • 71.Spivak B. Z., Andreev A. V., Magnetotransport phenomena related to the chiral anomaly in Weyl semimetals. Phys. Rev. B 93, 085107 (2016). [Google Scholar]
  • 72.Burkov A. A., Chiral anomaly and diffusive magnetotransport in Weyl metals. Phys. Rev. Lett. 113, 247203 (2014). [DOI] [PubMed] [Google Scholar]
  • 73.Ledwith P. J., Guo H., Levitov L., The hierarchy of excitation lifetimes in two-dimensional fermi gases. Ann. Phys. 411, 167913 (2019). [Google Scholar]
  • 74.Kiselev E. I., Schmalian J., Nonlocal hydrodynamic transport and collective excitations in Dirac fluids. Phys. Rev. B 102, 245434 (2020). [Google Scholar]
  • 75.Guo H., Ilseven E., Falkovich G., Levitov L. S., Higher-than-ballistic conduction of viscous electron flows. Proc. Natl. Acad. Sci. U.S.A. 114, 3068–3073 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Lucas A., Stokes paradox in electronic fermi liquids. Phys. Rev. B 95, 115425 (2017). [Google Scholar]
  • 77.Lucas A., Hartnoll S. A., Kinetic theory of transport for inhomogeneous electron fluids. Phys. Rev. B 97, 045105 (2018). [Google Scholar]
  • 78.Lamb H., On electrical motions in a spherical conductor. Philos. Trans. R. Soc. Lond. 174, 519–549 (1883). [Google Scholar]
  • 79.Reuter G. E. H., Sondheimer E. H., The theory of the anomalous skin effect in metals. Proc. R. Soc. Lond. A Math. Phys. Sci. 195, 336–364 (1948). [Google Scholar]
  • 80.Valentinis D., Optical signatures of shear collective modes in strongly interacting fermi liquids. Phys. Rev. Res. 3, 023076 (2021). [Google Scholar]
  • 81.Gurzhi R. N., Contribution to the theory of the skin effect in metals at low temperatures. Sov. Phys. JETP 20, 1228–1230 (1964). [Google Scholar]
  • 82.Duval C., Horváth Z., Horváthy P. A., Martina L., Stichel P. C., Berry phase correction to electron density in solids and “exotic” dynamics. Mod. Phys. Lett. B 20, 373–378 (2006). [DOI] [PubMed] [Google Scholar]
  • 83.Knoll A., Timm C., Meng T., Negative longitudinal magnetoconductance at weak fields in Weyl semimetals. Phys. Rev. B 101, 201402 (2020). [Google Scholar]
  • 84.Xiao C., et al., Linear magnetoresistance induced by intra-scattering semiclassics of Bloch electrons. Phys. Rev. B 101, 201410 (2020). [Google Scholar]
  • 85.Gao Y., Yang S. A., Niu Q., Geometrical effects in orbital magnetic susceptibility. Phys. Rev. B 91, 214405 (2015). [Google Scholar]
  • 86.Gao Y., Yang S. A., Niu Q., Field induced positional shift of Bloch electrons and its dynamical implications. Phys. Rev. Lett. 112, 166601 (2014). [DOI] [PubMed] [Google Scholar]
  • 87.Gorbar E. V., Miransky V. A., Shovkovy I. A., Sukhachov P. O., Second-order chiral kinetic theory: Chiral magnetic and pseudomagnetic waves. Phys. Rev. B 95, 205141 (2017). [DOI] [PubMed] [Google Scholar]
  • 88.Abbasi N., Taghinavaz F., Tavakol O., Magneto-transport in a chiral fluid from kinetic theory. J. High Energy Phys. 2019, 51 (2019). [Google Scholar]
  • 89.Yang S. Z., Gao J. H., Liang Z. T., Wang Q., Second-order charge currents and stress tensor in a chiral system. Phys. Rev. D 102, 116024 (2020). [Google Scholar]
  • 90.Ferziger J. H., Kaper H. G., Mathematical Theory of Transport Processes in Gases (North-Holland, 1972). [Google Scholar]
  • 91.Cercignani C., The Boltzmann Equation and Its Applications (Springer, 1988). [Google Scholar]
  • 92.Lucas A., Das Sarma S., Electronic sound modes and plasmons in hydrodynamic two-dimensional metals. Phys. Rev. B 97, 115449 (2018). [Google Scholar]
  • 93.Kumar N., et al., Extremely high magnetoresistance and conductivity in the type-II Weyl semimetals WP2 and MoP2. Nat. Commun. 8, 1642 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Abrikosov A. A., Fundamentals of the Theory of Metals (North-Holland, 1988). [Google Scholar]
  • 95.Sukhachov P. O., Glazman L. I., Anomalous electromagnetic field penetration in a Weyl or Dirac semimetal. arXiv [Preprint] (2021). https://arxiv.org/abs/2110.10167 (Accessed 25 January 2022). [DOI] [PubMed]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File
pnas.2200367119.sapp.pdf (339.7KB, pdf)

Data Availability Statement

There are no data underlying this work.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES