Hostname: page-component-7bb8b95d7b-pwrkn Total loading time: 0 Render date: 2024-09-30T03:16:38.827Z Has data issue: false hasContentIssue false

Electromagnetism of one-component plasmas of massless fermions

Published online by Cambridge University Press:  11 August 2023

V.M. Rylyuk*
Affiliation:
National Academy of Sciences of Ukraine “Center for Problems of Marine Geology, Geoecology and Sedimentary Ore Formation of the NAS of Ukraine”, Kyiv, 01054, Ukraine
I.M. Tkachenko
Affiliation:
Departament de Matemàtica Aplicada, Universitat Poliècnica de València, Valencia, 46022, Spain Al-Farabi Kazakh National University, Almaty, 050040, Kazakhstan
*
Email address for correspondence: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

We present a theoretical study of two- and three-dimensional massless Dirac one-component plasmas embedded in a constant uniform magnetic field. We determine the wavefunctions and Landau energy levels of a massless Dirac fermion in a constant magnetic field. On this basis we consider magnetism of Fermi fluids of massless charged particles. We show that such a three-dimensional Dirac plasma consisting of fermions with the same helicity has its own magnetic moment. We also consider the limit of strong magnetic fields and investigate the De Haas–van Alphen effect. We derive the Kubo formula for the electrical conductivity tensor of massless Dirac plasmas and consider the Shubnikov–de Haas effect. In addition, we propose a model of the static conductivity tensor and employ the matrix version of the classical method of moments to derive a Drude-like formula for the dynamic conductivity tensor for massless Dirac plasmas. We find that the electrical conductivity tensor for Dirac fermions with the right helicity is not isotropic in the plane perpendicular to the magnetic field.

Type
Research Article
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press

1. Introduction

In Sarma & Hwang (Reference Sarma and Hwang2009), the investigation of dynamic properties of massless Dirac one-component plasmas was carried out in one-dimensional (1-D), two-dimensional (2-D) and three-dimensional (3-D) cases at zero temperature in the random-phase approximation. In the present work we carry out a theoretical study of the electromagnetic properties of 3-D finite-temperature plasmas of massless fermions embedded in a constant uniform magnetic field. We define such a Dirac plasma as a system of charged carriers with the linear energy dispersion $\epsilon _{\boldsymbol {k}} = \hbar kv$ described by the Dirac–Weyl-like Hamiltonian

(1.1)\begin{equation} \hat{H}={-}{i}\hbar v\hat{\boldsymbol{\sigma}}\boldsymbol{\cdot}\boldsymbol{\nabla} . \end{equation}

Here $\hbar$ is the Planck constant, $\hat {\boldsymbol {\sigma }}$ is the set of Pauli matrices and $v$ is the constant speed of a Dirac fermion, which in (1.1) plays the role of the effective speed of light. Details regarding (1.1) are discussed in § 2. A very important new aspect of such a system is that the momentum $\hbar \boldsymbol {k}$ and the velocity magnitude $v$ decouple from each other: the momentum satisfies the equation of motion, but the velocity determines the displacement. The velocity magnitude remains constant and interactions change its direction only. In addition, the spin of a Dirac fermion can take only two directions: along the momentum or against it. In the first case, the fermion has the positive helicity, and in the second, the negative one.

The most common and studied case of a plasma of massless particles is the 2-D system of Dirac fermions in graphene (Novoselov et al. Reference Novoselov, Geim, Morozov, Jiang, Katsnelson, Grigorieva, Dubonos and Firsov2005; Katsnelson, Novoselov & Geim Reference Katsnelson, Novoselov and Geim2006; Castro Neto et al. Reference Castro Neto, Guinea, Peres, Novoselov and Geim2009), with the quasiparticles described by the Hamiltonian (1.1) as if they were massless relativistic particles with the momentum $\hbar \boldsymbol {k}$ (for example, photons) with the light speed $c$ substituted by the Fermi speed $v\approx c/300$. In Andersen, Jacobsen & Thygesen (Reference Andersen, Jacobsen and Thygesen2014) and Enaldiev (Reference Enaldiev2018), the existence of 1-D atomically confined plasmons at the edges of a zigzag nanoribbon was predicted, and collective excitations were considered in a two-component 1-D Dirac plasma. In Shaisultanov, Lyubarsky & Eichler (Reference Shaisultanov, Lyubarsky and Eichler2012), the 3-D ultrarelativistic plasma with the linear energy dispersion of quasiparticles was also studied in the context of quark–gluon plasmas (Heinz Reference Heinz2009; Moldabekov et al. Reference Moldabekov, Ludwig, Bonitz and Ramazanov2015). The spatial and magnetic confinement, the helical magnetic effect, and the chiral anomaly in Dirac plasmas were studied in Ren et al. (Reference Ren, Cheng, Li, Chao, Liu, Lv, Zhang, Sun and He2021) and Yamamoto & Yang (Reference Yamamoto and Yang2021).

In what follows we consider one-component 2-D and 3-D Dirac plasmas with a neutralizing background in a constant uniform magnetic field. The aim of this paper is to consider the electromagnetic properties of such a Dirac plasma. To this end, using the Dirac–Weyl equation, we determine the wavefunctions and Landau energy levels of a massless fermion in a constant magnetic field and demonstrate that these energy levels are formally similar to the energy spectrum of a 3-D relativistic electron in a constant magnetic field. However, unlike in the case of an electron with a non-zero mass, we find that the cyclotron frequency of massless fermions is explicitly non-classical being proportional to $1/\sqrt {\hbar }$. Further, we show that a 3-D Dirac plasma consisting of fermions with the same helicity in a weak magnetic field possesses its own magnetic moment, and we calculate this magnetic moment. In the case of strong magnetic fields we consider the De Haas–van Alphen (DHVA) effect, i.e. the giant oscillations of the Dirac plasma magnetization.

Next, within the framework of the linear response theory we derive a Kubo-like formula for the electrical conductivity tensor of a Dirac plasma. We apply it to consider the Shubnikov–de Haas (SdH) effect, the large amplitude oscillations of the 2-D and 3-D magnetized Dirac plasma conductivity. It stems from our study that both DHVA and SdH effects are direct consequences of the Landau diamagnetism. These results are complemented by numerical calculations of the magnetic moment and the static electrical conductivity of a Dirac plasma.

Finally, we apply the matrix version of the method of moments (see e.g. Kovalishina Reference Kovalishina1983; Adamyan & Tkachenko Reference Adamyan, Tkachenko and Verlag2000) to construct the Drude–Lorentz-like model of the Dirac plasma dynamic conductivity tensor. Additional aspects, namely, the normalization condition of the Fermi–Dirac distribution density in the cases of $2{\rm D}$ and 3-D massless fermions in a magnetic field and the mathematical details of the matrix method of moments are described in the Supplemental material available at https://doi.org/10.1017/S0022377823000752 and in appendix B.

The paper is organized as follows. Section 2 represents a brief overview of the theory of the Dirac and Weyl equations. In § 3 we solve the Dirac–Weyl equation for a massless Dirac fermion in a constant magnetic field. In § 4 we consider the magnetism of Dirac plasmas in weak magnetic fields and analyse the DHVA effect. Next, in § 5 we derive the Kubo formula for the electrical conductivity tensor of Dirac plasmas, consider the SdH effect in strong magnetic fields, and employ the classical method of moments to construct the dynamic internal conductivity tensor in the long-wavelength case. Section 6 concludes the work. In appendix A some coefficients are defined. In appendix B the matrix Hamburger problem, Nevanlinna's formula and the Drude–Lorentz-type formula for the conductivity tensor of a magnetized Dirac plasma are considered. Appendix C presents the static electrical conductivity tensor of a 3-D Dirac plasma in a magnetic field and in appendix D the normalization in a magnetic field is considered. In the Supplemental material the following are considered: (1) the Kramers–Kronig relations; (2) some mathematical formulae.

2. Dirac and Weyl equations

Relativistically invariant equations for electrons were obtained by Dirac. They consist of the following system of homogeneous first-order differential equations for two spinor functions $\hat {\varphi }$ and $\hat {\chi }$ (Berestetskii, Pitaevskii & Lifshitz Reference Berestetskii, Pitaevskii and Lifshitz1982):

(2.1a,b)\begin{equation} {i}\hbar \frac{\partial \hat{\varphi}}{\partial t} = mc^2\hat{\varphi}+c\hat{\boldsymbol{\sigma}}\boldsymbol{\cdot}\hat{\boldsymbol{p}} \hat{\chi},\quad {i}\hbar \frac{\partial \hat{\chi}}{\partial t}={-}mc^2\hat{\chi} + c\hat{\boldsymbol{\sigma}} \boldsymbol{\cdot}\hat{\boldsymbol{p}}\hat{\varphi} . \end{equation}

Here, $m$ is the electron mass, $\hat {\boldsymbol {p}}=-{\rm i}\hbar \boldsymbol {\nabla }$ is the momentum operator, $c$ is the speed of light and $\hat {\boldsymbol {\sigma }}$ is the set of Pauli matrices

(2.2a–c)\begin{equation} \hat{\sigma}_x = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, \quad \hat{\sigma}_y = \begin{pmatrix} 0 & -{\rm i} \\ {\rm i} & 0 \end{pmatrix}, \quad \hat{\sigma}_z = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix} . \end{equation}

In accordance to the general relation between the infinitesimal rotation operator and the angular momentum the matrix $\tfrac {1}{2}\hat {\boldsymbol {\sigma }}$ is the spin operator. The commutational properties of the Pauli matrices are those of the angular momentum operator components,

(2.3a–c)\begin{equation} \hat{\sigma}_x \hat{\sigma}_y - \hat{\sigma}_y \hat{\sigma}_x = 2{\rm i}\hat{\sigma}_z , \quad \hat{\sigma}_y \hat{\sigma}_z - \hat{\sigma}_z \hat{\sigma}_y = 2{\rm i}\hat{\sigma}_x , \quad \hat{\sigma}_z \hat{\sigma}_x - \hat{\sigma}_x \hat{\sigma}_z = 2{\rm i}\hat{\sigma}_y . \end{equation}

They satisfy also the following relations:

(2.4)\begin{equation} \hat{\sigma}_i \hat{\sigma}_k + \hat{\sigma}_k \hat{\sigma}_i = 2\delta_{ik} , \end{equation}

where $\delta _{ik}$ is the Kronecker delta. Let us introduce the bispinor function

(2.5)\begin{equation} \hat{\varPsi} = \begin{pmatrix} \hat{\varphi} \\ \hat{\chi} \end{pmatrix} . \end{equation}

The components of the bispinor (2.5) can be subjected to the unitary transformation

(2.6)\begin{equation} \hat{\varPsi}^{^{\prime }}= \begin{pmatrix} \hat{\xi} \\ \hat{\eta} \end{pmatrix} =\hat{\mathrm{U}}\hat{\varPsi}, \end{equation}

where $\hat {\mathrm {U}}$ is a certain unitary four-row matrix: $\hat {\mathrm {U}}\hat {\mathrm {U}}^{{\dagger} }=\hat {\mathrm {I}}$, for example

(2.7)\begin{equation} \hat{\mathrm{U}}=\frac{1}{\sqrt{2}} \begin{pmatrix} \hat{I} & \hat{I} \\ \hat{I} & -\hat{I} \end{pmatrix} , \end{equation}

which makes the spinors $\hat {\xi }$ and $\hat {\eta }$ satisfy the following equations:

(2.8a,b)\begin{equation} {i}\hbar \frac{\partial \hat{\xi}}{\partial t} = c\hat{\boldsymbol{\sigma}}\boldsymbol{\cdot}\hat{\boldsymbol{p}}\hat{\xi} + mc^{2} \hat{\eta} ,\quad {i}\hbar \frac{\partial \hat{\eta}}{\partial t} ={-} c\hat{\boldsymbol{\sigma}}\boldsymbol{\cdot}\hat{\boldsymbol{p}}\hat{\eta} + mc^{2}\hat{\xi} . \end{equation}

As it follows from (2.8a,b), due to the presence of the particle mass $m$, the first equation in (2.8a,b) for the spinor $\hat {\xi }$ contains the components of the spinor $\hat {\eta }$, while the second equation for $\hat {\eta }$ contains the components of the spinor $\hat {\xi }$. But if the particle mass is equal to zero ($m=0$), there is no such mixing of the components of the spinors, and the equations in (2.8a,b) transform into two independent ones,

(2.9)\begin{equation} {i}\hbar \frac{\partial}{\partial t} \begin{pmatrix} \hat{\xi} \\ \hat{\eta} \end{pmatrix} = c\hat{\boldsymbol{\sigma}}\boldsymbol{\cdot}\hat{\boldsymbol{p}} \begin{pmatrix} \hat{\xi} \\ -\hat{\eta} \end{pmatrix} \end{equation}

called the Weyl equations. These equations, like the original Dirac equations, are relativistically invariant. However, it can be easily seen that these equations will not be invariant under the transformation of space inversion. In other words, the two-component wavefunction satisfying (2.9), may be used to describe a particle of zero mass with the spin $1/2$ only if we drop the requirement of the invariance of the equations under space inversion. Such a situation exists in the case of the neutrinos which is described by (2.9); here the minus sign before $\hat {\eta }$ corresponds to the neutrino, while the plus sign corresponds to the antineutrino (Bjorken & Drell Reference Bjorken and Drell1964). The behaviour of the massless Dirac fermion in an external electromagnetic field with the vector potential $\boldsymbol {A}$ can also be described by the Weyl equations

(2.10)\begin{equation} {i}\hbar \frac{\partial}{\partial t} \begin{pmatrix} \hat{\xi} \\ \hat{\eta} \end{pmatrix} = v\hat{\boldsymbol{\sigma}}\boldsymbol{\cdot}\left(\hat{\boldsymbol{p}} + \frac{e}{c}\boldsymbol{A}\right) \begin{pmatrix} \hat{\xi} \\ -\hat{\eta} \end{pmatrix} , \end{equation}

with the speed of light $c$ replaced by the speed of a massless fermion $v$. The equation for the spinor $\hat {\xi }$ in (2.10) corresponds to the massless fermion with the right-handed helicity (the Dirac fermion spin is directed along the momentum $\boldsymbol {p}$) and the equation for the spinor $\hat {\eta }$ corresponds to the massless fermion with the left-handed helicity (with the spin directed against the momentum $\boldsymbol {p}$).

3. Massless fermions in a constant magnetic field

Let us determine the energy levels of a massless fermion in a constant uniform magnetic field. The vector potential of the uniform magnetic field $\boldsymbol {H}$ (with the $z$-axis directed along the magnetic field $\boldsymbol {H}$) is conveniently taken here as

(3.1a,b)\begin{equation} A_{x} ={-}Hy,\quad A_{y} = A_{z} = 0. \end{equation}

Then the equations in (2.10), in the 3-D case, can be written as

(3.2)\begin{equation} \left. \begin{aligned} {\rm i}\hbar \frac{\partial \hat{\xi}(\boldsymbol{r},t)}{\partial t} & = v\hat{\sigma}_x\left(\hat{p}_x - \frac{eH}{c}y\right)\hat{\xi}(\boldsymbol{r},t) + v\hat{\sigma}_y\hat{p}_y\hat{\xi}(\boldsymbol{r},t) + v\hat{\sigma}_z\hat{p}_z\hat{\xi}(\boldsymbol{r},t) , \\ {\rm i}\hbar \frac{\partial \hat{\eta}(\boldsymbol{r},t)}{\partial t} & ={-}v\hat{\sigma}_x\left(\hat{p}_x - \frac{eH}{c}y\right)\hat{\eta}(\boldsymbol{r},t) - v\hat{\sigma}_y\hat{p}_y\hat{\eta}(\boldsymbol{r},t) - v\hat{\sigma}_z\hat{p}_z\hat{\eta}(\boldsymbol{r},t) . \end{aligned} \right\} \end{equation}

First of all, we notice that the operators $\hat {p}_x$ and $\hat {p}_z$ commute with the right-hand sides of the equations in (3.2). This means that the $x$ and $z$ components of the generalized momentum are conserved. This is why we seek the solution of (3.2) in the form

(3.3)\begin{equation} \left. \begin{aligned} \hat{\xi}(\boldsymbol{r},t) & = \exp\left({\frac{{\rm i}}{\hbar}\left(p_x x + p_z z - \epsilon t\right)}\right) \hat{\xi}(y) , \quad \hat{\xi}(y) = \begin{pmatrix} u_+(y) \\ d_+(y)\end{pmatrix} , \\ \hat{\eta}(\boldsymbol{r},t) & = \exp\left({-\frac{{\rm i}}{\hbar}\left(p_x x + p_z z + \epsilon t\right)}\right) \hat{\eta}(y) , \quad \hat{\eta}(y) = \begin{pmatrix} u_{-}(y) \\ d_{-}(y)\end{pmatrix} , \end{aligned} \right\} \end{equation}

with the eigenvalues $p_x$ and $p_z$ being arbitrary real numbers. Since $A_z = 0$, the $z$-component of the generalized momentum can have any value, i.e. the motion of the massless Dirac fermion along the magnetic field is ‘not quantized’. In view of (3.3), the system of (3.2), transforms into $(\boldsymbol {p} = \hbar \boldsymbol {k})$ the following:

(3.4)\begin{equation} \left. \begin{aligned} v\hbar \frac{\partial u_+(\kern-0.01em y)}{\partial y} & = \left(\epsilon_+ + v \hbar k_z\right)d_+(\kern-0.01em y) + v \left(\frac{e}{c}H y - \hbar k_x\right)u_+(\kern-0.01em y) , \\ v\hbar \frac{\partial d_+(\kern-0.01em y)}{\partial y} & ={-}\left(\epsilon_+ - v \hbar k_z\right)u_+(\kern-0.01em y) - v\left(\frac{e}{c}H y - \hbar k_x\right)d_+(\kern-0.01em y) , \\ v\hbar \frac{\partial u_{-}(\kern-0.01em y)}{\partial y} & = \left(\epsilon_{-} - v \hbar k_z\right)d_{-}(\kern-0.01em y) - v \left(\frac{e}{c}H y - \hbar k_x\right)u_{-}(\kern-0.01em y) ,\\ v\hbar \frac{\partial d_{-}(\kern-0.01em y)}{\partial y} & ={-}\left(\epsilon_{-} + v \hbar k_z\right)u_{-}(\kern-0.01em y) + v\left(\frac{e}{c}H y - \hbar k_x\right)d_{-}(\kern-0.01em y) . \end{aligned} \right\} \end{equation}

Solving the system of differential equations in (3.4), after some algebra, we arrive at the following equations:

(3.5)\begin{equation} \frac{\partial^2 u_{n,\pm}(\zeta)}{\partial \zeta^2} + \frac{2}{\hbar^2 \omega_{{\rm H}}^2} \left\{\epsilon_{n,\pm}^2(k_z) - \frac{\hbar^2 \omega_{{\rm H}}^2}{2}\left[\left(\zeta - a_{{\rm H}} k_x\right)^2 \pm 1\right] - v^2\hbar^2 k_z^2\right\}u_{n,\pm}(\zeta) = 0 , \end{equation}

(with $\zeta =y/a_{\textrm {H}}$), which are formally similar to the equations for a linear oscillator with the frequency $\omega _{\textrm {H}} = v\sqrt {2eH/(c\hbar )} = \sqrt {2}v/a_{\textrm {H}}$, $a_{\textrm {H}} = \sqrt {c\hbar /(eH)}$ being the magnetic length and where

(3.6a,b)\begin{equation} \epsilon_{n,+}(k_z) = \hbar \sqrt{\omega_{{\rm H}}^2(n + 1) + v^2 k_z^2} , \quad \epsilon_{n,-}(k_z) = \hbar \sqrt{\omega_{{\rm H}}^2 n + v^2 k_z^2} \end{equation}

are the energy levels of a massless Dirac fermion in a constant uniform magnetic field with the right- ($\epsilon _{n,+}(k_z)$) and left-handed ($\epsilon _{n,-}(k_z)$) helicities. The eigenfunctions $u_{n,\pm }$ and $d_{n,\pm }$ corresponding to the energy levels (3.6a,b) are given by the expressions

(3.7)\begin{equation} \left. \begin{aligned} u_{n,\pm}(y) & = C_{n,\pm} \exp\left\{\frac{\left(y - y_0\right)^2}{2a_{{\rm H}}^2}\right\} H_n\left(\frac{y - y_0}{a_{{\rm H}}}\right) , \quad u_{{-}1,-}(y) = 0 , \quad y_0 = a_{{\rm H}}^2k_x , \\ d_{n,+}(y) & = 2\frac{\hbar v}{a_{{\rm H}}^2}\frac{a_{{\rm H}} n u_{n-1,+}(y) - (y - y_0)u_{n,+}(y)} {\epsilon_{n,+}(k_z) + v\hbar k_z} , \quad d_{n,-}(y) = 2\frac{\hbar v}{a_{{\rm H}}}\frac{n u_{n-1,-}(y)} {\epsilon_{n,-}(k_z) - v\hbar k_z} , \end{aligned} \right\} \end{equation}

where $y_{0}=a_{\textrm {H}}^2 k_x$, $H_n$ are the Hermite polynomials and the coefficients $C_{n,\pm }$ are given by (A1). In the 2-D case the third dimension is suppressed, i.e. in (3.2)–(3.7) and in (A1) we must put $k_z = 0$. In particular, the energy levels (3.6a,b) of a massless Dirac fermion in a constant magnetic field in the $2\textrm {D}$ case, are

(3.8a,b)\begin{equation} \epsilon_{n,+} = \hbar \omega_{{\rm H}}\sqrt{n + 1}, \quad \epsilon_{n,-} = \hbar \omega_{{\rm H}}\sqrt{n} . \end{equation}

Note that the expressions in (3.6a,b) for the energy levels are formally similar to the energy spectrum of a 3-D relativistic electron in a constant magnetic field (see e.g. Berestetskii et al. Reference Berestetskii, Pitaevskii and Lifshitz1982). However, the cyclotron frequency $\omega _{\mathrm {H}}$ of a massless fermion, not like that of an electron, is proportional to the square root $\sqrt {H}$ and increases with the velocity $v$ of the Dirac fermion. Besides, it must be emphasized, that the cyclotron frequency of a massless fermion is explicitly non-classical being proportional to $1/\sqrt {\hbar }$ (McCann & Fal'ko Reference McCann and Fal'ko2006). In addition, as follows from (A1), we do not have explicit expressions for the coefficients $C_{n,\pm }$, but only the recurrent relations.

Since the energy levels in (3.6a,b) do not contain an arbitrary real $k_x$, this energy spectrum is continuously degenerate with the finite degeneracy degree for the motion in the $xy$-plane restricted to a large, but finite, area $S=L_x L_y$, and with the number of states (for given $n$) being $eHS/(2{\rm \pi} \hbar c)$. Then the 2-D number of states is

(3.9)\begin{equation} \frac{eHS}{2{\rm \pi} \hbar c} . \end{equation}

If, in addition, the motion is restricted in the $z$-direction also (dimension $L_z$), we observe that the number of possible values of $k_z$ in an interval $\Delta k_z$ is $(L_z/2{\rm \pi} ) \Delta k_z$ and the 3-Dnumber of states in this interval reads

(3.10)\begin{equation} \frac{eHS}{2{\rm \pi} \hbar c}\frac{L_z}{2{\rm \pi}}\Delta k_z = \frac{eHV}{4{\rm \pi}^2 \hbar c}\Delta k_z . \end{equation}

4. Magnetism of Dirac plasmas

4.1. Weak magnetic fields

Each Dirac fermion has a spin and in the presence of an external magnetic field the spins of these fermions are polarized and a net magnetic moment arises (the Pauli paramagnetism). Besides, when a gas consisting of massless fermions is placed in a magnetic field, the motion of the fermions changes: they begin to move along the helical trajectories. Hence, these magnetized Dirac fermions generate an additional magnetic field, whose direction is opposite to that of the external field. This is the Landau diamagnetism effect due to the variation of the orbital motion of fermions caused by the external field. This effect is of a quantum nature and is absent in the classical approximation. In fact, the diamagnetism occurs and results from the quantization of massless fermions levels in the magnetic field (see § 3). We conclude that the magnetization of a Dirac plasma of massless fermions results from the Pauli paramagnetism and the Landau diamagnetism.

Let us calculate the thermodynamic potential (Landau & Lifshitz Reference Landau and Lifshitz1980)

(4.1)\begin{equation} \varOmega ={-}\beta^{{-}1}\sum_n \ln \left[1 + \exp \left(\eta - \beta \epsilon_n\right)\right] , \end{equation}

where $\beta ^{-1} = T$ is the system temperature in energy units (throughout the article we put the Boltzmann constant $k_{\textrm {B}} = 1$), and the sum is taken over all possible states, in the case of weak magnetic fields, i.e. when the inequality $\beta \hbar \omega _{\textrm {H}} \ll 1$ holds. The dimensionless chemical potential $\eta$ in (4.1), in 3-D plasmas, is determined by the condition (D2). Introducing the density of states in accordance with (3.10), we obtain

(4.2)\begin{equation} \varOmega_{{\pm}} ={-}\frac{V}{2{\rm \pi}^2 \hbar}\frac{eH}{\beta c}\int_0^{\infty}\,{d}\epsilon \sum_{n}\frac{{d}k_{z}}{{d}\epsilon}\ln \left\{1 + \exp \left[\eta -\beta \epsilon_{n,\pm}(k_z) \right]\right\} . \end{equation}

At small $H$ we may replace the summation by integration using the Euler–Maclaurin sum formula (2.1) in the Supplemental material. In the case of a degenerate Dirac plasma ($\beta \epsilon _{\textrm {F}} \gg 1$), for the right helicity ($+$), we put in equation (2.1) in the Supplemental material $a = 0$, $b = \epsilon _{\textrm {F}}^2/(\hbar ^2 \omega _{\textrm {H}}^2) - 1$, and for the left helicity ($-$) we apply $a = 1$, $b = \epsilon _{\textrm {F}}^2/(\hbar ^2 \omega _{\textrm {H}}^2)$, where $\epsilon _{\textrm {F}} = \hbar v k_{\textrm {F}}$ is the Fermi energy with $k_{\textrm {F}} = \sqrt [3]{3{\rm \pi} ^2 n_3}$ ($n_3$ is the 3-D Dirac fermion density). As a result, we obtain

(4.3a–c)\begin{equation} \varOmega_{{\pm}} ={-}\frac{V}{8{\rm \pi}^2}\frac{\epsilon_{{\rm F}}^2 \omega_{{\rm H}}^2}{\hbar v^3}S_{{\pm}} , \quad S_+{=}\sum_{n=0}^{[n_{0,+}]}f_+(n) ,\quad S_{-} = \sum_{n=1}^{[n_{0,-}]} \left[1 + f_{-}(n)\right] , \end{equation}

where $[x]$ is the integer part of number $x$,

(4.4)\begin{equation} \left. \begin{aligned} f_+(n) & = \sqrt{\frac{n_{0,+} - n}{n_{0,+} + 1}} - \frac{n+1}{n_{0,+} + 1} \ln\left(\sqrt{\frac{n_{0,+} + 1}{n+1}} + \sqrt{\frac{n_{0,+} - n}{n+1}}\right) , \\ f_{-}(n) & = \sqrt{1 - \frac{n}{n_{0,-}}} - \frac{n}{n_{0,-}} \ln\left(\sqrt{\frac{n_{0,-}}{n}} + \sqrt{\frac{n_{0,-}}{n} - 1}\right) \end{aligned} \right\} \end{equation}

and

(4.5a,b)\begin{equation} n_{0,+} = \epsilon_{{\rm F}}^2/(\hbar^2 \omega_{{\rm H}}^2) - 1 , \quad n_{0,-} = \epsilon_{{\rm F}}^2/(\hbar^2 \omega_{{\rm H}}^2) . \end{equation}

The magnetic moment per unit volume is defined as

(4.6)\begin{equation} \boldsymbol{\mathfrak{M}}={-}\frac{1}{V}\left(\frac{\partial \varOmega}{\partial \boldsymbol{H}}\right)_{\beta ,V,\eta}. \end{equation}

From (4.3ac) and (4.4), using the Euler–Maclaurin formula (2.1) in the Supplemental material, after some algebra, in the limit $H \rightarrow \boldsymbol {0}$ we arrive at the expression for the 3-D magnetic moments

(4.7)\begin{equation} \boldsymbol{\mathfrak{M}}_{{\rm F},\pm}\simeq{\mp} \boldsymbol{\mathfrak{m}}_{{\rm F}} + \frac{5\alpha}{48{\rm \pi}^2} \frac{v}{c}\left[1-\frac{4}{5}\ln \left(\frac{\hbar \omega_{{\rm H}}}{2\epsilon_{F}}\right)\right] \boldsymbol{H}, \end{equation}

where $\alpha = e^2/(\hbar c)\simeq 1/137$ is the fine structure constant and

(4.8)\begin{equation} |\boldsymbol{\mathfrak{m}}_{{\rm F}}|=\frac{e}{8{\rm \pi}^2}\frac{\epsilon_{{\rm F}}^2}{\hbar^2 vc} \end{equation}

is the proper magnetic moment of the degenerate Dirac plasma. Then the total magnetic moment of the degenerate massless Dirac plasma is

(4.9)\begin{equation} \boldsymbol{\mathfrak{M}}_{{\rm F}, {\rm tot}} = \boldsymbol{\mathfrak{M}}_{{\rm F},+} + \boldsymbol{\mathfrak{M}}_{{\rm F},-}\simeq \frac{5\alpha}{24{\rm \pi}^2}\frac{v}{c}\left[1-\frac{4}{5} \ln \left(\frac{\hbar \omega_{{\rm H}}}{2\epsilon_{{\rm F}}}\right)\right]\boldsymbol{H} . \end{equation}

In addition, for the (linear) magnetic susceptibility of degenerate Dirac plasmas we obtain

(4.10)\begin{equation} \chi_{{\rm F},\pm} \simeq \frac{\alpha}{16{\rm \pi}^2}\frac{v}{c} \left[1-\frac{4}{3}\ln \left(\frac{\hbar \omega_{{\rm H}}}{2\epsilon_{\rm F}}\right)\right] \end{equation}

and

(4.11)\begin{equation} \chi_{{\rm F}, {\rm tot}}\simeq \frac{\alpha}{8{\rm \pi}^2}\frac{v}{c}\left[1 - \frac{4}{3} \ln \left(\frac{\hbar \omega_{\rm H}}{2\epsilon_{\rm F}}\right)\right] . \end{equation}

In the opposite case of high temperatures ($\beta \epsilon _\textrm {F}\ll 1$) the Dirac fermions form a Boltzmann gas, and the magnetic moments are

(4.12)\begin{equation} \boldsymbol{\mathfrak{M}}_{{\rm B},\pm} \simeq{\mp} \boldsymbol{\mathfrak{m}}_{\rm B} + \frac{5ne}{64} \frac{\omega_{\rm H}}{c}l^{2}\left(1 - \frac{4}{15}\beta \hbar \omega_{\rm H}\right) \frac{\boldsymbol{H}}{H}, \end{equation}

where

(4.13)\begin{equation} |\boldsymbol{\mathfrak{m}}_{\rm B}|=\frac{lne}{8}\frac{v}{c} \end{equation}

is the proper magnetic moment of a Dirac plasma in the case of the Boltzmann statistics. The total magnetic moment of a Boltzmann massless Dirac plasma reads

(4.14)\begin{equation} \boldsymbol{\mathfrak{M}}_{{\rm B}, {\rm tot}} \simeq \frac{5ne}{32}\frac{\omega_{\rm H}}{c}l^2 \left(1-\frac{4}{15}\beta \hbar \omega_{\rm H}\right)\frac{\boldsymbol{H}}{H}. \end{equation}

In turn, for the magnetic susceptibilities we have

(4.15)\begin{equation} \chi_{{\rm B},\pm }\simeq \frac{15}{64{\rm \pi} \beta }\frac{l^{2}\omega_{p}^{2}}{\hbar \omega_{\rm H}c^{2}} \left(1-\frac{8}{15}\beta \hbar \omega_{\rm H}\right) \end{equation}

and

(4.16)\begin{equation} \chi_{{\rm B}, {\rm tot}}\simeq \frac{15}{32{\rm \pi} \beta }\frac{l^{2}\omega_{p}^{2}} {\hbar \omega_{\rm H}c^{2}}\left(1 - \frac{8}{15}\beta \hbar \omega_{\rm H}\right) , \end{equation}

where $l = \beta \hbar v$, $\omega _p = \sqrt {4{\rm \pi} n_3 e^2/\mu }$ is the plasma frequency and $\mu = 3/(\beta v^2 )$ is the effective Dirac ‘mass’. As it stems from (4.7) and (4.12), the massless plasma, consisting of fermions with only positive or only negative helicity, possesses the proper magnetic moment in the absence of an external magnetic field. In other words, it possesses its own magnetization.

4.2. Strong magnetic fields

In the overdense plasma (e.g. at the critical electron number density $n_e \sim 10^{21}$ $\textrm {cm}^{-3}$), under the influence a laser pulse with the wavelength $\sim 1.054\,\mathrm {\mu }\textrm {m}$, of duration $\tau _0 \sim 1$ ps and with a high-intensity $\sim 10^{20}\,\textrm {W}\,\textrm {cm}^{-2}$, were generated ultrastrong magnetic fields in the range 50–700 MG. In these experiments (Tatarakis et al. Reference Tatarakis, Watts, Beg, Clark, Dangor, Gopal, Haines, Norreys, Wagner, Wei, Zepf and Krushelnick2002a,Reference Tatarakis, Gopal, Watts, Beg, Wei, Dangor, Krushelnick, Wagner, Norreys, Clark, Zepf and Evans2002b; Wagner et al. Reference Wagner, Tatarakis, Gopal, Beg, Clark, Dangor, Evans, Haines, Mangles and Norreys2004; Sarri et al. Reference Sarri, Macchi, Cecchetti, Kar, Liseykina, Yang, Dieckmann, Fuchs, Galimberti and Gizzi2012) atoms in the laser plasma are ionized in the tunnelling limit ($\gamma \ll 1$) for a time $\tau \sim 1$ fs ($\tau \ll \tau _0$), in the barrier suppression regime.

Let us now consider strong magnetic fields under the condition $1\lesssim \beta \hbar \omega _\textrm {H} \ll |\eta |$. In this case the magnetization of the Dirac plasma contains a contribution which oscillates with a large amplitude as a function of $\boldsymbol {H}$, DHVA effect (Landau & Lifshitz Reference Landau and Lifshitz1980). To separate the oscillatory parts of the thermodynamic potential (4.1) it is convenient to transform the sum in (4.2) by means of the Poisson formula (2.2) in the Supplemental material. In the 3-D case this leads to

(4.17)\begin{gather} \varOmega_{{\pm}} = \varOmega_{0,\pm}(\mu) + \frac{V}{{\rm \pi}^{2}\hbar}\frac{eH}{\beta c}{\rm Re} \sum_{n=1}^{\infty }I_{n}, \end{gather}
(4.18)\begin{gather} I_n ={-}\int_{-\infty}^{\infty}\int_0^{\infty}\ln \left[1+\exp \left(\eta -\beta \epsilon_{x,\pm}(k_z)\right)\right]\exp \left(2{\rm \pi} {\rm i}nx\right)\,{\rm d}x\,{\rm d}k_z, \end{gather}

where $\varOmega _{0,\pm }(\mu )$ is the thermodynamic potential in the absence of the field. In the integrals $I_n$ the variable $x$ is taken to be $x = (\epsilon ^2 - \hbar ^2\omega _\textrm {H}^2 - v^2\hbar ^2 k_z^2)/ (\hbar ^2\omega _\textrm {H}^2)$ for the right-handed helicity and $x = (\epsilon ^2 - v^2\hbar ^2 k_z^2)/ (\hbar ^2\omega _\textrm {H}^2)$ for the left-handed helicity. Then, for the required oscillatory part of the integrals we have

(4.19)\begin{equation} I_{\text{osc}, n} ={-}\int_{-\infty }^{\infty}\int_{0}^{\infty}\ln \left[1 + \exp \left(\eta - \beta \epsilon \right)\right]\exp \left(\frac{2{\rm \pi} in\epsilon^2} {\hbar^2 \omega_{\rm H}^2}\right)\exp \left(-\frac{2{\rm \pi} in v^2 k_z^2}{\omega_{\rm H}^2}\right) \,{\rm d}\epsilon \,{\rm d}k_z . \end{equation}

The main contribution to the integral over $k_z$ proceeds from the values of $k_z \sim \omega _\textrm {H}/v$ (Landau & Lifshitz Reference Landau and Lifshitz1980). On the other hand, the oscillatory part of the integral is produced by the values of $\epsilon$ near $\eta /\beta$ (see below); the lower limit of integration over $\epsilon$ is therefore taken to be equal to zero. The integration over $k_z$ is separable and is carried out by means of the Poisson integral (2.4) in the Supplemental material. The remaining integral is taken by parts and as a result we obtain

(4.20)\begin{equation} \varOmega_{\text{osc},\pm} = \frac{\sqrt{2}V}{16{\rm \pi}^{3}\beta}\frac{\omega_{\rm H}^{3}}{v^{3}} \sum_{n=1}^{\infty}\frac{1}{n^{3/2}}\int_{0}^{\infty}\frac{\cos \left(\dfrac{2{\rm \pi} y^{2}n} {\beta^2\hbar^2\omega_{\rm H}^2} + \dfrac{\rm \pi}{4}\right)}{\exp(\kern-0.01em y - \eta ) + 1}\,{{\rm d} y}, \end{equation}

where the dimensionless chemical potential $\eta$ is determined by the condition (D2). If the inequality holds $\eta \gg 1$, we make in this integral the change of variable $y - \eta = \xi$, and using the value of the complex integral $J_1$ in equation (2.5) in the Supplemental material, from (4.20), we finally get for the oscillatory part of $\varOmega _{\pm }$,

(4.21)\begin{equation} \varOmega_{\text{osc},\pm}\simeq \dfrac{\sqrt{2}V}{16{\rm \pi}^2\beta} \dfrac{\omega_{\rm H}^3}{v^3}\sum_{n=1}^{\infty}\dfrac{\sin \left(\dfrac{2{\rm \pi} \eta^2 n} {\beta^2\hbar^2\omega_{\rm H}^2} + \dfrac{\rm \pi}{4}\right)}{n^{3/2} \sinh \left[4{\rm \pi}^2n\eta /(\beta^2\hbar^2\omega_{\rm H}^2)\right]}. \end{equation}

In the case if $\eta \lesssim 1$, the integral in (4.20) can be calculated only numerically.

The main contribution to the magnetic moment comes from the differentiation of the most rapidly varying factors in (4.21), i.e. the sines in the numerators. This leads to

(4.22)\begin{equation} \mathfrak{M}_{{\rm osc}, \pm} \simeq \dfrac{e}{2\sqrt{2}{\rm \pi}}\dfrac{\eta^{2}V}{\beta^{3}\hbar^{3} \omega_{\rm H} vc}\sum_{n=1}^{\infty}\dfrac{\cos \left(\dfrac{2{\rm \pi} \eta^{2}n}{\beta^{2}\hbar^{2} \omega_{\rm H}^{2}} + \dfrac{\rm \pi}{4}\right)}{\sqrt{n}\sinh \left[4{\rm \pi}^{2}n\eta /(\beta^{2}\hbar^{2} \omega_{\rm H}^{2})\right]}. \end{equation}

For 2-D Dirac plasmas, in view of equation (2.2) in the Supplemental material,

(4.23)\begin{equation} \varOmega_{{\pm}} = \varOmega_{0,\pm}(\mu) + \dfrac{2S}{\rm \pi}\dfrac{eH}{\beta c}{\rm Re} \sum_{n=1}^{\infty}I_n \end{equation}

and the oscillatory parts of the thermodynamic potential are

(4.24)\begin{equation} I_{\text{osc},n} ={-}\int_0^{\infty}\ln \left[1 + \exp \left(\eta - \beta \epsilon \right) \right]\exp \left(\frac{2{\rm \pi} in\epsilon^2}{\hbar^2 \omega_{\rm H}^2}\right)\,{\rm d}\epsilon . \end{equation}

The integral in (4.24) is taken by parts and then

(4.25)\begin{equation} \varOmega_{\text{osc},\pm} = \dfrac{\hbar S}{4{\rm \pi}^2\beta}\dfrac{\omega_{\rm H}^2}{v^2} \sum_{n=1}^{\infty}\dfrac{1}{n}\int_0^{\infty}\dfrac{\cos \left(\dfrac{2{\rm \pi} y^2 n} {\beta^2\hbar^2\omega_{\rm H}^2}\right)}{\exp(y - \eta) + 1}\,{{\rm d} y}, \end{equation}

where the dimensionless chemical potential $\eta$ is determined by the condition (D12). If the inequality holds $\eta \gg 1$, we make in this integral the change of variable $y-\eta = \xi$, and in view of equation (2.5) in the Supplemental material, we obtain

(4.26)\begin{equation} \varOmega_{\text{osc},\pm}\simeq \dfrac{\hbar S}{4{\rm \pi}^2\beta}\dfrac{\omega_{\rm H}^2}{v^2} \sum_{n=1}^{\infty}\dfrac{\sin \left(\dfrac{2{\rm \pi} \eta^2 n}{\beta^2\hbar^2\omega_{\rm H}^2}\right)} {n\sinh \left[4{\rm \pi}^2 n\eta /(\beta^2\hbar^2\omega_{\rm H}^2)\right]} . \end{equation}

Finally, this leads to the following result for the oscillatory part of the magnetic moment in the 2-D case:

(4.27)\begin{equation} \mathfrak{M}_{{\rm osc},\pm} \simeq \dfrac{e\eta^{2}S}{\beta^3\hbar^2\omega_{\rm H}^2 c} \sum_{n=1}^{\infty}\dfrac{\cos \left(\dfrac{2{\rm \pi} \eta^2 n}{\beta^2\hbar^2\omega_{\rm H}^2}\right)} {\sinh \left[4{\rm \pi}^2 n\eta /(\beta^2\hbar^2\omega_{\rm H}^2)\right]}. \end{equation}

The functions in (4.22) and (4.27) oscillate with high frequency, which is a manifestation of the DHVA effect in 3-D and 2-D massless Dirac plasmas. Its ‘period’ in the variable $1/H$ is

(4.28)\begin{equation} \Delta \left(1/H\right) = 2e\dfrac{\beta^2\hbar v^2}{c\eta^2} \end{equation}

and it increases with the Dirac fermion speed as $v^{2}$. The dimensionless chemical potential $\eta$ in (4.28) is determined by the conditions (D2) and (D12). In the case of non-degenerate hot Dirac plasmas we have constructed analytical approximations (D10) and (D18) for the dimensionless chemical potential $\eta$. As it can be seen from the above calculations the DHVA effect is a direct consequence of the Landau diamagnetism.

Some results of numerical calculations for the value of the normalized magnetic moments $\mathfrak {M}^{\ast }$ (3-D) and $\mathfrak {M}_\textrm {2D}$ (2-D) of Dirac plasmas versus the magnetic field $H$ for various densities and temperatures are presented in figures 1 and 2. The magnitude of the magnetic field $H$ is measured in megagauss (MG), and $n_3$ and $n_2$ are the 3-D and 2-D Dirac fermion density, respectively. The dimensionless chemical potential $\eta$ for these figures is determined by solving (D2) and (D12) numerically. These figures clearly demonstrate the DHVA effect in magnetized massless 3-D and 2-D Dirac plasmas.

Figure 1. The function $\mathfrak {M}^{*}$ versus $H$ for $n_3 = 10^{22}$ cm$^{-3}$, $T = 10$ K and $v \approx c/300$.

Figure 2. The function $\mathfrak {M}_\textrm {2D}$ versus $H$ for $n_2 = 10^{18}$ cm$^{-2}$, $T = 100$ K and $v \approx c/300$.

5. Conductivity tensor of Dirac plasmas in an external constant magnetic field

5.1. The Kubo linear reaction theory

Let a system consisting of interacting massless fermions be in a state of statistical equilibrium, which is described by the Gibbs statistical operator (Akhiezer & Peletminskii)

(5.1)\begin{equation} \hat{w} = \exp \{\varOmega -\beta \hat{\mathcal{H}}_0 - \eta \hat{N})\} , \end{equation}

where $\hat {\mathcal {H}}_0$ is the Hamiltonian of interacting Dirac fermions and $\hat {N}$ is the particle number operator. The Gibbs operator (5.1) acts in the Hilbert space of vectors describing the quantum states of massless fermions in a magnetic field. At some time $t_0$ the external field is switched on, so that the Hamiltonian operator of the system becomes

(5.2)\begin{equation} \hat{\mathcal{H}}(t) = \hat{\mathcal{H}}_0 + \hat{V}^{\rm ext}(t), \end{equation}

where

(5.3)\begin{equation} \hat{V}^{\rm ext}(t) ={-}\dfrac{1}{c}\int \,{d}\boldsymbol{r}\boldsymbol{A}^{\rm ext}(\boldsymbol{r},t)\boldsymbol{\cdot}\hat{\boldsymbol{j}}(\boldsymbol{r}) \end{equation}

is the interaction Hamiltonian in the Weyl gauge with an external electromagnetic field characterized by the vector potential $\boldsymbol {A}^\textrm {ext}(\boldsymbol {r},t)$ and the current-density operator $\hat {\boldsymbol {j}}(\boldsymbol {r})$ is determined by (5.25a,b). The statistical operator $\hat {\rho }(t)$ of the system of Dirac fermions satisfies the Liouville equation

(5.4)\begin{equation} {\rm i}\hbar \dfrac{\partial \hat{\rho}(t)}{\partial t}=\left[\hat{\mathcal{H}}(t), \hat{\rho}(t)\right] , \end{equation}

where $[\hat {A},\hat {B}]$ is the commutator of the operators $\hat {A}$ and $\hat {B}$. In order to find the statistical operator of the system $\hat {\rho }(t)$ for $t > t_0$, we introduce the operator

(5.5)\begin{equation} \tilde{\hat{\rho}}(t) = \exp({{\rm i}\hat{\mathcal{H}}_0 t/\hbar})\hat{\rho}(t) \exp({-{\rm i}\hat{\mathcal{H}}_0 t/\hbar}) \end{equation}

in the interaction picture. Then according to (5.4) we obtain for $\tilde {\hat {\rho }}(t)$ the following equation:

(5.6a,b)\begin{equation} {\rm i}\hbar \dfrac{\partial \tilde{\hat{\rho}}(t)}{\partial t} = \left[\tilde{\hat{V}}^{\rm ext}(t),\tilde{\hat{\rho}}(t)\right] ,\quad \tilde{\hat{V}}^{\rm ext}(t) = \exp({{\rm i}\hat{\mathcal{H}}_0 t/\hbar})\hat{V}^{\rm ext}(t) \exp({-{\rm i}\hat{\mathcal{H}}_0 t/\hbar}) . \end{equation}

We assume that for $t=-\infty$ there was no external field and the system was in a statistical equilibrium state, i.e. $\hat {\rho }(-\infty )=\hat {w}$. Since $[\hat {\mathcal {H}}_0,\hat {w}] = 0$, then $\tilde {\rho }(-\infty )=\hat {w}$. The last equality is the initial condition for (5.6a,b). Then from (5.6a,b) we obtain the integral equation for the operator $\tilde {\rho }(t)$,

(5.7)\begin{equation} \tilde{\hat{\rho}}(t) = \hat{w}-\dfrac{{\rm i}}{\hbar}\int_{-\infty}^{t}\,{\rm d}t^{^{\prime}} \left[\tilde{\hat{V}}^{\rm ext}(t^{^{\prime}}), \tilde{\hat{\rho}}(t^{^{\prime}})\right]. \end{equation}

In the linear approximation for the interaction with the external field, we have

(5.8)\begin{equation} \tilde{\hat{\rho}}(t)\simeq \hat{w}-\dfrac{{\rm i}}{\hbar}\int_{-\infty}^{t}\,{\rm d}t^{^{\prime}} \left[\tilde{\hat{V}}^{\rm ext}(t^{^{\prime}}), \hat{w}\right] . \end{equation}

Let us define the average value of the induced current-density in the system of massless fermions as

(5.9)\begin{equation} \bar{\boldsymbol{j}}(\boldsymbol{r},t) = \mathrm{Tr}\left\{\hat{\rho}(t)\hat{\boldsymbol{j}}(\boldsymbol{r})\right\} . \end{equation}

Taking into account that in a statistical equilibrium the average current is equal to zero, from (5.8) we obtain

(5.10)\begin{equation} \bar{\boldsymbol{j}}(\boldsymbol{r},t) = \dfrac{{\rm i}}{\hbar}\int_{-\infty}^{t}\,{\rm d}t^{^{\prime}}{\rm Tr} \left\{\hat{w}\left[\tilde{\hat{V}}^{\rm ext}(t^{^{\prime}}), \tilde{\hat{\boldsymbol{j}}}(\boldsymbol{r},t)\right]\right\} , \end{equation}

where

(5.11)\begin{equation} \tilde{\hat{\boldsymbol{j}}}(\boldsymbol{r},t) = \exp({{\rm i}\hat{\mathcal{H}}_0 t/\hbar})\hat{\boldsymbol{j}}(\boldsymbol{r}) \exp({-{\rm i}\hat{\mathcal{H}}_0 t/\hbar}) . \end{equation}

In view of (5.3), the $\mu$ – component of the average current is

(5.12)\begin{equation} \overline{j_{\mu}}(\boldsymbol{r},t) ={-}\dfrac{1}{c}\int_{-\infty}^{\infty}\,{\rm d}t^{^{\prime}} \int \,{\rm d}\boldsymbol{r}^{^{\prime}}G_{\mu \nu }(\boldsymbol{r}-\boldsymbol{r}^{^{\prime}}, t-t^{^{\prime}}) A_{\nu}^{\rm ext}(\boldsymbol{r}^{^{\prime}},t^{^{\prime}}) , \end{equation}

where the Green function current–current $G_{\mu \nu }$ is defined as

(5.13)\begin{equation} G_{\mu \nu}(\boldsymbol{r},t) ={-}\dfrac{{\rm i}}{\hbar}\theta (t){\rm Tr} \left\{w\left[j_{\mu}(\boldsymbol{r},t), j_{\nu}(0)\right]\right\} , \end{equation}

where $\theta (t)$ is the Heaviside step function. Passing in (5.12) to the Fourier components according to equation (2.6) in the Supplemental material, we have

(5.14)\begin{equation} \overline{j_{\mu}}(\boldsymbol{k},\omega) ={-}\frac{1}{c}G_{\mu \nu}(\boldsymbol{k},\omega) A_{\nu}^{\rm ext}(\boldsymbol{k},\omega) . \end{equation}

Taking into account that at $\varphi (\boldsymbol {r},t) = 0$

(5.15)\begin{equation} \boldsymbol{E}^{\rm ext}(\boldsymbol{r},t) ={-}\frac{1}{c}\frac{\partial \boldsymbol{A}^{\rm ext}(\boldsymbol{r},t)}{\partial t} , \end{equation}

or for the Fourier components,

(5.16)\begin{equation} \boldsymbol{E}^{\rm ext}(\boldsymbol{k},\omega) = {\rm i}\frac{\omega}{c}\boldsymbol{A}^{\rm ext}(\boldsymbol{k}, \omega), \end{equation}

(5.14) can be rewritten as

(5.17)\begin{equation} \overline{j_{\mu}}(\boldsymbol{k},\omega) = \frac{{\rm i}}{\omega}G_{\mu \nu}(\boldsymbol{k},\omega) E_{\nu}^{\rm ext}(\boldsymbol{k},\omega). \end{equation}

Then, according to the definition

(5.18)\begin{equation} \overline{j_{\mu}}(\boldsymbol{k},\omega) = \kappa_{\mu \nu}(\boldsymbol{k},\omega) E_{\nu}^{\rm ext}(\boldsymbol{k},\omega), \end{equation}

the electrical susceptibility tensor of the system of massless Dirac fermions can be expressed in terms of the Green function as follows:

(5.19)\begin{equation} \kappa_{\mu \nu}(\boldsymbol{k},\omega) = \frac{{\rm i}}{\omega}G_{\mu \nu}(\boldsymbol{k},\omega) , \end{equation}

or, in view of (5.13),

(5.20)\begin{equation} \kappa_{\mu \nu}(\boldsymbol{k}, \omega) = \frac{1}{\hbar \omega V}\int_{0}^{\infty} \exp({-\varepsilon t + {\rm i}\omega t})\left\langle\left[\hat{j}_{\boldsymbol{k}, \mu}(t), \hat{j}_{-\boldsymbol{k}, \nu}(0)\right]\right\rangle\,{\rm d}t \quad (\varepsilon > 0) , \end{equation}

where $\langle \hat {f}(t)\rangle \equiv \textrm {Tr}(\hat {w}\hat {f}(t))$. In equation (5.20) introduced is an infinitesimally small positive quantity $\varepsilon$ for adiabatic switching on of a periodic external electric field. The limit $\varepsilon \downarrow 0$ is taken after the thermodynamic limit $V \rightarrow \infty$ ($V/N = \textrm {const.}$), $N$ is the number of plasma particles.

5.2. Electrical conductivity tensor

Oscillations similar to those of the DHVA effect are observed in kinetic phenomena, for example, in the electrical conductivity. Here we will consider conductivity oscillations of a Dirac plasma embedded in a constant strong magnetic field, i.e. the SdH effect (see e.g. Landau & Lifshitz Reference Landau and Lifshitz1980), using a static conductivity model within the framework of the Kubo linear response theory.

Along with the electrical susceptibility tensor in (5.18), the conductivity tensor $\hat {\sigma }(\boldsymbol {k},\omega )$ can also be defined, as a reaction to the total electromagnetic field in the plasma. In the absence of the spatial dispersion, i.e. at $\boldsymbol {k}\rightarrow \boldsymbol {0}$, for the conductivity tensor $\hat {\sigma }(\omega ) = \lim _{\boldsymbol {k}\rightarrow \boldsymbol {0}}\hat {\sigma }(\boldsymbol {k},\omega )$ the Kubo formula (5.20) is valid (Zubarev Reference Zubarev1974),

(5.21)\begin{equation} \sigma_{\mu \nu}(\omega) = \frac{1}{\hbar \omega V}\lim_{\varepsilon \downarrow 0} \lim_{\boldsymbol{k}\rightarrow \boldsymbol{0}}\int_{0}^{\infty} \exp({-\varepsilon t + {\rm i}\omega t})\left\langle \left[\hat{j}_{\boldsymbol{k},\mu}(t),\hat{j}_{-\boldsymbol{k},\nu}(0)\right]\right\rangle \,{\rm d}t . \end{equation}

Then, the Hermitian part of the electrical conductivity tensor (5.21) of a Dirac plasma in a magnetic field reads

(5.22)\begin{equation} \sigma_{\mu \nu ,{\rm H}}(\omega) = \sigma_{\mu \nu ,{\rm H},+}(\omega) + \sigma_{\mu \nu ,{\rm H},-}(\omega) , \end{equation}

where

(5.23)\begin{equation} \sigma_{\mu \nu ,{\rm H},\pm}(\omega) = \frac{\rm \pi}{\hbar \omega}\sum_{n,m} \left[f_{\rm FD}(\epsilon_{n,\pm}) - f_{\rm FD}(\epsilon_{n,\pm} + \hbar \omega)\right] \left\langle n|j_{\mu}|m\right\rangle_{{\pm}}\left\langle m|j_{\nu}|n \right\rangle_{{\pm}}\delta (\omega - \omega_{m,n,\pm}) \end{equation}

are the conductivity tensors corresponding to right-handed ($+$) and left-handed ($-$) helicities, the subscript $H$ denotes the Hermitian matrix,

(5.24)\begin{equation} f_{\rm FD}(\epsilon_{n,\pm}) = \frac{1}{\exp(\beta \epsilon_{n,\pm} - \eta) + 1} \end{equation}

is the Fermi–Dirac distribution density, which obeys the normalization condition (D2) for the 3-D case and (D12) for the 2-D case, and $\omega _{n,m,\pm } = (\epsilon _{n,\pm } - \epsilon _{m,\pm })/\hbar$. In equation (5.23)

(5.25a,b)\begin{equation} \left\langle n|j_{\mu}|m\right\rangle_{{\pm}} = \int \,{\rm d}\boldsymbol{r} \left\langle n|j_{\mu}(\boldsymbol{r})|m\right\rangle_{{\pm}} , \quad \hat{\boldsymbol{j}}_+{=} -ev\hat{\xi}^+\hat{\boldsymbol{\sigma}}\hat{\xi}, \quad \hat{\boldsymbol{j}}_{-} = ev\hat{\eta}^+\hat{\boldsymbol{\sigma}}\hat{\eta} \end{equation}

are the matrix elements for the right-handed and the left-handed helicities, in the bra-ket (Dirac) notation, of the current-density operator $\boldsymbol {j}(\boldsymbol {r})$.

To account for the scattering of the Dirac fermions in the relaxation time approximation, the $\delta$-function with the energy conservation law in (5.23) should be ‘smeared out’ by replacing the adiabatic parameter $\Delta$ with the collision frequency $\nu _{\pm }$ (Skobov & Kaner Reference Skobov and Kaner1964; Blank & Kaner Reference Blank and Kaner1966). Then

(5.26)\begin{align} \sigma_{\mu \nu, {\rm H},\pm}(\omega) & = \sum_{n,m}\frac{f_{\rm FD}(\epsilon_{n,\pm}) - f_{\rm FD}(\epsilon_{n,\pm} + \hbar \omega)}{\hbar \omega} \nonumber\\ & \quad \times \frac{\nu_{{\pm}}}{(\omega - \omega_{m,n,\pm})^2 + \nu_{{\pm}}^2} \langle n|j_{\mu}|m\rangle_{{\pm}}\langle m|j_{\nu}|n\rangle_{{\pm}} \end{align}

and in the static limit $\sigma _{0,\mu \nu,\pm } = \lim _{\omega \rightarrow 0} \sigma _{\mu \nu,\textrm {H},\pm }(\omega )$, we obtain

(5.27)\begin{equation} \sigma _{0,\mu \nu ,\pm} ={-}\frac{eH}{4{\rm \pi}^2 \hbar c}\int_{-\infty}^{\infty} \sum_{n,m=0}^{\infty}\left. \frac{\partial f_{\rm FD}(\epsilon ,k_{z})}{\partial \epsilon} \right\vert_{\epsilon = \epsilon_{n,\pm}}\frac{\nu_{{\pm}}}{\omega_{n,m,\pm}^{2} + \nu_{{\pm}}^2}\left\langle n|j_{\mu}|m\right\rangle_{{\pm}} \left\langle m|j_{\nu}|n\right\rangle_{{\pm}}\,{\rm d}k_z. \end{equation}

The relaxation time approximation in (5.26) is equivalent to replacing the collision integral in the kinetic equation with the quantity $\sim \delta f /\tau _{\pm }$ (Abrikosov Reference Abrikosov1988), where $\delta f$ is the small correction to the Fermi–Dirac distribution due to collisions, and the relaxation time $\tau _{\pm } = 1/\nu _{\pm }$. The collision frequency $\nu _{\pm }$ is calculated in (5.34).

Using the Dirichlet formula (2.7) in the Supplemental material the expression in (5.27) for the static electrical conductivity tensor can be transformed into

(5.28)\begin{equation} \sigma_{0,\mu \nu ,\pm} ={-}\frac{eH}{4{\rm \pi} ^{2}\hbar c}\int_{-\infty}^{\infty} \sum_{m=0}^{\infty}\sum_{n=0}^{\infty}\left\{\mathcal{F}_{n,n+m,\pm}(k_{z}) + \mathcal{F}_{n+m,n,\pm}(k_{z})\right\}\,{\rm d}k_z, \end{equation}

where $\mathcal {F}_{n,n+m,\pm }(k_{z})$ and $\mathcal {F}_{n+m,n,\pm }(k_{z})$ are given by equation (2.8) in the Supplemental material.

Taking into account of (3.7), after some algebra, the Hermitian part of the static electrical conductivity tensor of a 3-D Dirac plasma in a magnetic field can be represented as

(5.29)\begin{equation} \hat{\sigma}_0 = \hat{\sigma}_{0, +} + \hat{\sigma}_{0, -} , \quad \hat{\sigma}_{0,\pm} = \begin{pmatrix} \sigma _{xx,\pm} & {\rm i}\sigma_{xy,\pm} & 0 \\ -{\rm i}\sigma_{xy,\pm} & \sigma_{yy,\pm} & 0 \\ 0 & 0 & \sigma_{zz,\pm} \end{pmatrix} , \end{equation}

where the Cartesian components $\sigma _{ij,\pm }$ ($i,j = x,y,z$) are given by (C1).

In turn, the static electrical conductivity tensor of a 2-D magnetized massless Dirac plasma takes the form

(5.30)\begin{equation} \hat{\sigma}_0 = \hat{\sigma}_{0,+} + \hat{\sigma}_{0,-} , \quad \hat{\sigma}_{0,\pm} = \begin{pmatrix} \sigma_{xx,\pm} & {\rm i}\sigma_{xy,\pm} \\ -{\rm i}\sigma_{xy,\pm} & \sigma_{yy,\pm} \\ \end{pmatrix} , \end{equation}

where

(5.31)\begin{equation} \left. \begin{aligned} \sigma_{xx, \pm} & ={-}2{\rm \pi} \omega_{\rm H}^2\frac{\hbar^4 v^2}{R_k}\sum_{n = 0}^{\infty} \left[\mathfrak{F}(\epsilon_{n,\pm}) + \mathfrak{F}(\epsilon_{n+1,\pm})\right] \mathcal{J}_{n,n+1,\pm}(x,x,0) , \\ \sigma_{yy, \pm} & ={-}2{\rm \pi} \omega_{\rm H}^2\frac{\hbar^4 v^2}{R_k}\sum_{n = 0}^{\infty} \left[\mathfrak{F}(\epsilon_{n,\pm}) + \mathfrak{F}(\epsilon_{n+1,\pm})\right] \mathcal{J}_{n,n+1,\pm}(y,y,0) , \\ \sigma_{xy, \pm} & ={-}2{\rm \pi} \omega_{\rm H}^2\frac{\hbar^4 v^2}{R_k}\sum_{n = 0}^{\infty} \left[\mathfrak{F}(\epsilon_{n,\pm}) - \mathfrak{F}(\epsilon_{n+1,\pm})\right] \mathcal{J}_{n,n+1,\pm}(x,y,0) \end{aligned} \right\} \end{equation}

and $\epsilon _{n,\pm }$ are given by (3.8a,b). In particular, for a degenerate Dirac plasma ($\beta \epsilon _\textrm {F} \gg 1$), $\mathfrak {F}(\epsilon _{n,\pm }) = - \delta (\epsilon _{n,\pm } - \epsilon _\textrm {F})$ and the 2-D electrical conductivity tensor (5.31) is

(5.32)\begin{equation} \left. \begin{aligned} \sigma_{ii,-} & = 4{\rm \pi} \epsilon_{\rm F} \frac{\hbar^2 v^2}{R_k} \left[\mathcal{J}_{[n_0],[n_0]+1,-}(i,i,0) + \mathcal{J}_{[n_0]-1,[n_0],-}(i,i,0)\right] ,\\ \sigma_{ii,+} & = 4{\rm \pi} \epsilon_{\rm F} \frac{\hbar^2 v^2}{R_k} \left[\mathcal{J}_{[n_0]-1,[n_0],+}(i,i,0) + \mathcal{J}_{[n_0]-2,[n_0]-1,+}(i,i,0)\right] , \quad i = x, y\\ \sigma_{xy,-} & = 4{\rm \pi} \epsilon_{\rm F} \frac{\hbar^2 v^2}{R_k} \left[\mathcal{J}_{[n_0],[n_0]+1,-}(x,y,0) - \mathcal{J}_{[n_0]-1,[n_0],-}(x,y,0)\right] ,\\ \sigma_{xy,+} & = 4{\rm \pi} \epsilon_{\rm F} \frac{\hbar^2 v^2}{R_k} \left[\mathcal{J}_{[n_0]-1,[n_0],+}(x,y,0) - \mathcal{J}_{[n_0]-2,[n_0]-1,+}(x,y,0)\right] , \end{aligned} \right\} \end{equation}

where $[n_0]$ is the integer part of $n_0 = \epsilon _\textrm {F}^2/(\hbar ^2 \omega _\textrm {H}^2)$. For example, for the surface number density of Dirac fermions $n_s = 10^{21}$ cm$^{-2}$ and the magnetic field $H = 30$ MG, $[n_0] = 1$. Then, taking into account that $\omega _\textrm {H} \gg \nu _{\pm }$, we obtain

(5.33a–c)\begin{equation} \sigma_{xx} \simeq \frac{\nu}{\omega_{\rm H}}\frac{59}{R_k} , \quad \sigma_{yy} \simeq \frac{\nu}{\omega_{\rm H}}\frac{61}{R_k} , \quad \sigma_{xy} \simeq \frac{\nu}{\omega_{\rm H}}\frac{57}{R_k} , \end{equation}

where $\nu \simeq \nu _{-} \simeq \nu _+$ and $\sigma _{ij} = \sigma _{ij,-} + \sigma _{ij,+}$ ($i,j = x,y$).

Note, that the conductivity of a ‘plasma consisting of charges with a non-zero mass’ is not necessarily isotropic in the plane perpendicular to the magnetic field. This is a consequence of the fact that the currents $\hat {j}_x$ and $\hat {j}_y$ in (5.25a,b) are expressed in terms of the different Pauli matrices $\hat {\sigma }_x$ and $\hat {\sigma }_y$. However, as follows from (A2), $J_{x,n,-}(k_z) = J_{y,n,-}(k_z)$ and therefore $\sigma _{xx,-} = \sigma _{yy,-}$. In other words, for fermions with negative helicity, the static electrical conductivity is isotropic in the plane perpendicular to the magnetic field. But as it can be seen from (A2), $J_{x,n,+}(k_z)\neq J_{y,n,+}(k_z)$ and consequently for fermions with positive helicity $\sigma _{xx,+} \neq \sigma _{yy,+}$.

To calculate the collision frequency $\nu _{\pm }$ in (C1) and (C2a,b), we shall consider scattering from impurities and for the sake of simplicity we assume it to be isotropic. Within the first Born approximation the collision frequency in the presence of a magnetic field is (Abrikosov Reference Abrikosov1988)

(5.34)\begin{equation} \nu_{{\pm}} = \frac{\omega_{\rm H}^2}{4{\rm \pi}}\frac{n_i|v_i|^2}{\hbar^2 v^2}J_{{\pm}} , \quad J_{{\pm}} ={-}\hbar \sum_n \int \left. \frac{\partial f_{\rm FD}(\epsilon)}{\partial \epsilon} \right|_{\epsilon = \epsilon_{n,\pm}}\,{\rm d}k_z = J_{0,\pm} + J_{{\rm osc},\pm} , \end{equation}

where $n_i$ is the impurity atom number density and $v_i$ is the interaction potential of an electron with an impurity atom. According to Poisson's formula (2.2) in the Supplemental material the sum $J_{0,\pm }$ in (5.34) is given by equation (2.3) in the Supplemental material. In turn, for the oscillatory part of the integral, in the 3-D case, we arrive at the following expression:

(5.35)\begin{equation} J_{{\rm osc},\pm} = \frac{4\beta}{\hbar\omega_{\rm H}^2}{\rm Re}\sum_{n=1}^{\infty}J_{n} , \end{equation}

where

(5.36)\begin{equation} J_n = \int_{-\infty}^{\infty}\int_0^{\infty}\frac{\exp \left(\beta\epsilon - \eta \right)} {\left[\exp \left(\beta \epsilon -\eta \right) + 1\right]^2} \exp \left(\frac{2{\rm \pi} in\epsilon^2}{\hbar^2\omega_{\rm H}^2}\right) \exp \left(-\frac{2{\rm \pi} inv^2 k_z^2}{\omega_{\rm H}^2}\right)\epsilon \,{\rm d}\epsilon \,{\rm d}k_z . \end{equation}

The integration over $k_z$ in (5.36) is carried out using equation (2.4) in the Supplemental material and we obtain

(5.37)\begin{equation} J_{{\rm osc},\pm} = \frac{\beta}{\sqrt{2}\hbar v\omega_{\rm H}}\sum_{n=1}^{\infty}\frac{1}{\sqrt{n}} \int_0^{\infty}\frac{\exp \left(\beta \epsilon -\eta \right)} {\left[\exp \left(\beta \epsilon -\eta \right) + 1\right]^2} \sin \left(\frac{2{\rm \pi} n\epsilon^2}{\hbar^{2}\omega_{\rm H}^2} + \frac{\rm \pi}{4}\right) \epsilon \,{\rm d}\epsilon , \end{equation}

where the dimensionless chemical potential $\eta$ is determined by the condition (D2). If the inequality holds $\eta \gg 1$, we make in this integral the change of variable $y-\eta = \xi$ and using the value of the integral $J_2$ in equation (2.5) in the Supplemental material, from (5.37), we finally have for the oscillatory part

(5.38)\begin{equation} J_{\mathrm{osc},\pm}\simeq 8\sqrt{2}{\rm \pi}^2\dfrac{\eta^2}{\beta^3 \hbar^3\omega_{\rm H}^{3}v} \sum_{n=1}^{\infty }\dfrac{\sqrt{n}\sin \left(\dfrac{2{\rm \pi} \eta^2 n}{\beta^{2}\hbar^{2}\omega_{\rm H}^{2}} +\dfrac{\rm \pi}{4}\right)}{\sinh \left[4{\rm \pi}^2 n\eta /(\beta^{2}\hbar^{2}\omega_{\rm H}^{2})\right]} . \end{equation}

In the case if $\eta \lesssim 1$, the integral in (5.37) can be calculated only numerically.

For 2-D Dirac plasmas, the integral in (5.36) can be written as

(5.39)\begin{equation} J_n = \int_0^{\infty}\dfrac{\exp \left(\beta\epsilon - \eta \right)} {\left[\exp \left(\beta \epsilon -\eta \right) + 1\right]^2} \exp \left(\dfrac{2{\rm \pi} in\epsilon^2}{\hbar^2 \omega_{\rm H}^2}\right)\epsilon \,{\rm d}\epsilon \end{equation}

and $J_{\mathrm {osc},\pm }$ in (5.37) transforms into

(5.40)\begin{equation} J_{\mathrm{osc},\pm} = \dfrac{\beta}{\hbar \omega_{\rm H}^2}\sum_{n=1}^{\infty}\int_0^{\infty} \dfrac{\exp \left(\beta \epsilon -\eta \right)} {\left[\exp \left(\beta \epsilon -\eta \right) + 1\right]^2} \sin \left(\dfrac{2{\rm \pi} n\epsilon^2}{\hbar^2\omega_{\rm H}^2}\right)\epsilon \,{\rm d}\epsilon , \end{equation}

where the dimensionless chemical potential $\eta$ is determined by the condition (D12). If the inequality holds $\eta \gg 1$, we make in this integral the change of variable $y-\eta = \xi$, and in view of the complex integral $J_2$ in equation (2.5) in the Supplemental material, we finally obtain

(5.41)\begin{equation} J_{\mathrm{osc},\pm}\simeq 16{\rm \pi}^2\dfrac{\eta^2}{\beta^3 \hbar^3 \omega_{\rm H}^4} \sum_{n=1}^{\infty}\dfrac{n\sin \left(\dfrac{2{\rm \pi} \eta^2 n}{\beta^2 \hbar^2 \omega_{\rm H}^2}\right)} {\sinh \left[4{\rm \pi}^2 n \eta /(\beta^2 \hbar^2 \omega_{\rm H}^2)\right]} . \end{equation}

The functions in (5.38) and (5.41) oscillate with high frequency, which is a manifestation of the SdH effect in 3-D and 2-D massless Dirac plasmas. Its ‘period’ in the variable $1/H$ is the same as in DHVA effect,

(5.42)\begin{equation} \Delta \left(1/H\right) = 2\mathrm{e}\dfrac{\beta^2 \hbar v^2}{c\eta^2} \end{equation}

and it increases with the Dirac fermion speed as $v^{2}$. The dimensionless chemical potential $\eta$ in (5.42) is determined by the conditions (D2) and (D12). As it can be seen from the above calculations, the SdH effect is a direct consequence of Landau diamagnetism.

Some results of numerical calculations for the functions $\sigma = \sigma _{xx,\pm }, \sigma _{yy,\pm }$, $R_{zz} = (\sigma _{zz,\pm })^{-1}$ for the 3-D massless Dirac plasma, and for $\sigma _\textrm {2D} = \sigma _{xx,\pm }, \sigma _{yy,\pm }$ for the 2-D Dirac plasma as a function of the magnetic field $H$ for various densities and temperatures are shown in figures 3–5. Arbitrary units are used for the conductivity, the magnitude of the magnetic field $H$ is measured in MG, and $n_3$ and $n_2$ are the 3-D and 2-D Dirac fermion density, respectively. The dimensionless chemical potential $\eta$ for these figures is determined by solving (D2) and (D12) numerically. These figures clearly demonstrate the SdH effect in magnetized Dirac plasmas.

Figure 3. The function $\sigma$ versus $H$ for $n_3 = 10^{21}$ cm$^{-3}$, $T = 10$ K and $v \approx c/300$.

Figure 4. The function $R_{zz}$ versus $H$ for $n_3 = 10^{21}$ cm$^{-3}$, $T = 100$ K and $v \approx c/300$.

Figure 5. The function $\sigma _\textrm {2D}$ versus $H$ for $n_2 = 10^{19}$ cm$^{-2}$, $T = 10$ K and $v \approx c/300$.

5.3. Dynamic conductivity tensor. The Drude–Lorentz form

As it is stressed in the Supplemental material (see § 1), in the long-wavelength case both the electrical susceptibility and the internal conductivity are response functions. With $\sigma _\textrm {H}(\omega )$ being the positive-definite Hermitian part (5.23)

(5.43)\begin{equation} \hat{\sigma}_{\rm H}(\omega) = \dfrac{1}{2}\left[\hat{\sigma}^{{\dagger}}(\omega) + \hat{\sigma}(\omega)\right] \end{equation}

of the conductivity tensor $\boldsymbol {\hat {\sigma }}(\omega )$ of a Dirac plasma, we introduce its power frequency moments,

(5.44)\begin{equation} {\hat{{\rm T}}}_{\ell}=\dfrac{1}{\rm \pi}\int_{-\infty}^{\infty}\omega^{\ell}\hat{\sigma}_{\rm H}(\omega) \,{\rm d}\omega , \quad \ell = 0,1 . \end{equation}

The moments (5.44) can be calculated using (5.23), as follows:

(5.45)\begin{equation} \left. \begin{aligned} {\rm T}_{0,\mu \nu} & = \dfrac{1}{\hbar}\sum_{n,m}\dfrac{f_{\rm FD}(\epsilon_{n,\pm}) - f_{\rm FD}(\epsilon_{n,\pm} + \hbar \omega_{m,n,\pm})}{\omega_{m,n,\pm}}\left\langle n| j_{\mu}|m\right\rangle_{{\pm}}\left\langle m|j_{\nu}|n\right\rangle_{{\pm}} ,\\ {\rm T}_{1,\mu \nu} & =\dfrac{1}{\hbar}\sum_{n,m}\left[f_{\rm FD}(\epsilon_{n,\pm}) - f_{\rm FD}(\epsilon_{n,\pm} + \hbar \omega_{m,n,\pm})\right] \left\langle n|j_{\mu}|m\right\rangle_{{\pm}}\left\langle m|j_{\nu}|n\right\rangle_{{\pm}} . \end{aligned} \right\} \end{equation}

We wish to reconstruct the tensor $\hat {\sigma }(z)$ for $\textrm {Im}\ z \geq 0$ as a non-canonical solution of the truncated matrix Hamburger moment problem involving only two moments (5.45). Using the matrix version of the Nevanlinna theorem (Kovalishina (Reference Kovalishina1983), Adamyan & Tkachenko (Reference Adamyan, Tkachenko and Verlag2000) and also see appendix B), we arrive at

(5.46)\begin{equation} \hat{\sigma}(z) = \hat{\sigma}_0\left({\hat{{\rm I}}} + {\rm i}{\hat{\varOmega}}_1\hat{\tau} \right)\left[{\hat{{\rm I}}}-{\rm i}\left(z{\hat{{\rm I}}}-{\hat{\varOmega}}_1\right)\hat{\tau}\right]^{{-}1} , \end{equation}

where

(5.47)\begin{equation} {\hat{\varOmega}}_1 = {\hat{{\rm T}}}_0^{{-}1}{\hat{{\rm T}}}_1 . \end{equation}

The dynamic conductivity tensor (5.46) is expressed in terms of the static conductivity tensor $\hat {\sigma }_0$ in (5.29) or (5.30) and the relaxation time tensor $\hat {\tau } = \hat {\nu }^{-1}$ ($\hat {\nu }$ is the collision frequency tensor). In the simplest approximation we can put $\hat {\nu } = {\hat {\textrm {I}}} \nu _{\pm }$, where $\nu _{\pm }$ is given by (5.34). Thus, the dynamic conductivity tensor will also oscillate with high frequency, which is a manifestation of the SdH effect in magnetized massless Dirac plasmas.

The dielectric tensor can be constructed and the spectrum of collective excitations in the system under scrutiny can be studied, but this problem is beyond the scope of the present paper.

6. Conclusions

We have considered 2-D and 3-D one-component massless Dirac plasma embedded in a constant uniform magnetic field. First of all we have determined the wavefunctions and the energy Landau levels of the massless Dirac fermion in a constant magnetic field. Our solution is formally similar to the wavefunctions and energy spectrum of the relativistic electron in a constant magnetic field, but with the cyclotron frequency proportional to the velocity $v$ of the massless Dirac fermion and the square root of the magnetic field $\sqrt {H}$. We also found that the cyclotron frequency of the Dirac fermion is explicitly non-classical, being proportional to $1/\sqrt {\hbar }$.

Next, we investigated the magnetism. We calculated the magnetic moment for weak magnetic fields and showed that a 3-D massless Dirac plasma consisting of fermions with the same helicity has its own magnetic moment even in the absence of a magnetic field. In the opposite limit of strong magnetic fields we have considered the DHVA effect – oscillations of the Dirac plasma magnetization with a large amplitude. In the case if the inequality $\eta = \mu /T\gg 1$ holds, the oscillatory part of the magnetic moment can be evaluated analytically and the period of these oscillations decreases with increasing speed of massless Dirac fermions.

Within the framework of the linear response theory we derived the Kubo formula for the electrical conductivity tensor. We proposed a model of the 3-D and 2-D static electrical conductivity tensor and used the matrix version of the classical method of moments to derive a Drude-like formula for the dynamic conductivity tensor of magnetized massless Dirac plasmas. Note that the case of 2-D plasma can have important applications in nanoelectronics. We find that the electrical conductivity tensor for 3-D and 2-D massless Dirac plasmas, in the absence of the spatial dispersion, is not isotropic in the plane perpendicular to the magnetic field. We have shown that both static and dynamic conductivities are characterized by high frequency oscillations which is a manifestation of the SdH effect in magnetized plasmas. The period of these oscillations decreases with increasing speed of massless Dirac fermions. Note, that oscillations of the conductivity are most convenient for experimental observations. It must be emphasized that both DHVA and SdH effects are a direct consequence of Landau diamagnetism.

To support our theory we performed numerical calculations of the magnetic moment and the static conductivity. Our figures clearly demonstrate peculiarities of DHVA and SdH effects in 3-D and 2-D magnetized massless Dirac plasmas. Besides, we have considered the question of a general connection between the external and internal formalisms in the electrodynamics of anisotropic homogeneous media, little reviewed in theliterature.

Supplementary material

Supplementary material is available at https://doi.org/10.1017/S0022377823000752 for the Kramers–Kronig relations and some mathematical formulae.

Acknowledgements

We are deeply grateful to A.L. Mitler for performing numerical calculations.

Editor A. Schekochihin thanks the referees for their advice in evaluating this article.

Funding

I.M.T. acknowledges the support provided by the Ministry of Education and Science, Kazakhstan via grant #AP09260349.

Declaration of interests

The authors report no conflict of interest.

Appendix A. The coefficients $C_{n,\pm }$ and $J_{i,n,\pm }(k_z)$

The coefficients $C_{n,\pm }$ in (3.7) are defined as

(A1)\begin{equation} \left. \begin{aligned} C_{n,+} & = \dfrac{b_n C_{n-1,+} + \sqrt{a_n + b_n \left(b_n - a_n\right)C_{n-1,+}^2}}{a_n} , \\ C_{0,+} & = \dfrac{1}{\sqrt{a_{\rm H} \sqrt{\rm \pi}}\sqrt{1 + \dfrac{2\hbar^2 v^2} {a_{\rm H}^2 \left[\epsilon_{0,+}(k_z) + v\hbar k_z\right]^2}}} ,\\ a_n & = a_{\rm H} \sqrt{\rm \pi}2^nn!\left\{1 + \dfrac{2\hbar^2 v^2(2n+1)} {a_{\rm H}^2 \left[\epsilon_{n,+}(k_z) + v\hbar k_z \right]^2}\right\} ,\\ b_n & = \dfrac{\sqrt{\rm \pi}\hbar^2 v^2 2^{n+1}nn!}{a_{\rm H}\left[\epsilon_{n,+}(k_z) + v\hbar k_z\right]^2} ,\\ C_{n,-} & = \sqrt{\dfrac{1}{a_{\rm H}\sqrt{\rm \pi}2^n n!} - \dfrac{2n\hbar^2 v^2} {a_{\rm H}^2 \left[\epsilon_{n,-}(k_z) - \hbar v k_z\right]^2}\ C_{n-1,-}^2} , \\ C_{0,-} & = \dfrac{1}{{\rm \pi}^{1/4}\sqrt{a_{\rm H}}}. \end{aligned} \right\} \end{equation}

The functions $J_{i,n,\pm }(k_z)$ in (C1) and (C2a,b) are defined by the following relations:

(A2)\begin{equation} \left. \begin{aligned} J_{x,n,+}(k_z) & = 2^n(n+1)! \left[\left(\frac{1}{\epsilon_{n+1,+}(k_z) + v\hbar k_z} + \frac{1}{\epsilon_{n,+}(k_z) + v\hbar k_z}\right)C_{n+1,+} \right. \\ & \quad \left. - \frac{C_{n,+}}{\epsilon_{n+1,+}(k_z) + v\hbar k_z}\right]C_{n,+} ,\\ J_{y,n,+}(k_z) & ={-}2^n(n+1)! \left[\left(\frac{1}{\epsilon_{n+1,+}(k_z) + v\hbar k_z} + \frac{1} {\epsilon_{n,+}(k_z) + v\hbar k_z}\right)C_{n+1,+} \right. \\ & \quad \left. + \frac{C_{n,+}}{\epsilon_{n+1,+}(k_z) + v\hbar k_z}\right]C_{n,+} ,\\ J_{z,n,2,+}(k_z) & = 2^n(n+2)!\frac{\left(C_{n+2,+} - C_{n+1,+}\right)C_{n,+}} {\left(\epsilon_{n,+}(k_z) + v\hbar k_z\right)\left(\epsilon_{n+2,+}(k_z) + v\hbar k_z\right)} , \\ J_{x,n,-}(k_z) & = J_{y,n,-}(k_z) = 2^n(n+1)!\frac{C_{n-1,-}C_{n+1,-}}{\epsilon_{n+1,-}(k_z) - v\hbar k_z} , \\ J_{z,n,0,+}(k_z) & = 1 - \sqrt{\rm \pi}a2^{n+1}n!C_{n,+}^2 ,\\ J_{z,n,0,-}(k_z) & = 1 - \sqrt{\rm \pi}a2^{n+1}n!C_{n,-}^2 . \end{aligned} \right\} \end{equation}

Appendix B. The matrix Hamburger problem and Nevanlinna's formula

Here we consider the mathematical details of the matrix truncated Hamburger moment problem. The truncated Hamburger problem for matrix moments is formulated in the following way (Adamyan & Tkachenko (Reference Adamyan, Tkachenko and Verlag2000); see also Akhiezer Reference Akhiezer1965): Given a set of Hermitian $m\times m$ matrices

(B1)\begin{equation} \left\{ \hat{\mathrm{T}}_{0},\hat{\mathrm{T}}_{1},\hat{\mathrm{T}}_{2},\ldots,\hat{\mathrm{T}}_{2\mathrm{n}}\right\} ,\quad n=0,1,2,\ldots , \end{equation}

to find all matrix measures $\hat {\boldsymbol {\sigma }}(t)$ such that

(B2)\begin{equation} \int_{-\infty }^{+\infty }t^{k}\,{\rm d}\hat{\boldsymbol{\sigma}}(t)=\hat{\mathrm{T}_{k}} ,\quad k=0,1,2,\ldots ,2n. \end{equation}

In this paper we consider the case with $n=1$, i.e. with only three known matrix moments $\{ \hat {\mathrm {T}}_{0},\hat {\mathrm {T}}_{1},\hat {\mathrm {T}}_{2}\}$. In other words, the aim is to construct a tensorial generalization of the Drude–Lorentz formula. The Nevanlinna theorem establishes a bijection between the solutions of the truncated Hamburger matrix problem of moments $\{\hat {\mathrm {T}}_{0},\hat {\mathrm {T}}_{1},\hat {\mathrm {T}}_{2}\}$ and the Nevanlinna parameter tensor function $\hat {\mathrm {R}}(z)$ which is also a response or Nevanlinna class function (Krein & Nudel'man Reference Krein and Nudel'man1977) such that

(B3)\begin{equation} \lim_{z\rightarrow \infty}\left(\hat{\mathrm{R}}(z)/z\right) = 0 ,\quad {\rm Im}\ z>0. \end{equation}

The coefficients of this linear-fractional transformation are the matrix orthogonal polynomials and their conjugates constructed as it is shortly described below (Adamyan & Tkachenko Reference Adamyan, Tkachenko and Verlag2000).

Consider the invertible Hankel block matrices

(B4)\begin{equation} \hat{\varGamma}_{0}=\hat{\mathrm{T}}_{0},\quad \hat{\varGamma}_{1}=\left( \begin{array}{cc} \hat{\mathrm{T}}_{0} & \hat{\mathrm{T}}_{1} \\ \hat{\mathrm{T}}_{1} & \hat{\mathrm{T}}_{2} \end{array} \right) ,\ldots \end{equation}

and the matrices

(B5)\begin{equation} \left. \begin{aligned} \hat{\varXi}_{0} & =\hat{\mathrm{I}},\quad \hat{\varTheta}_{0}(t)=\hat{\mathrm{I}},\\ \hat{\boldsymbol{\varXi}}_{1} & =\left( 0,\hat{\mathrm{I}}\right) ^{T},\quad \hat{ \boldsymbol{\varTheta}}_{1}(t)=\left( \hat{\mathrm{I}},t\hat{\mathrm{I}}\right) , \end{aligned} \right\} \end{equation}

where $\hat {\mathrm {I}}$ is the $2\times 2$ unit matrix. Let us introduce $2\times 2$ matrix polynomials

(B6)\begin{equation} \left. \begin{aligned} \hat{\mathrm{D}}_{s}(t) & =\hat{\boldsymbol{\varTheta}}_{s}(t)\hat{\varGamma} _{s}^{{-}1}\hat{\boldsymbol{\varXi}}_{s} ,\quad \hat{\mathrm{D}}_{0}(t)=\hat{\varTheta}_{0}(t) \hat{\varGamma}_{0}^{{-}1}\hat{\varXi}_{0}=\hat{\mathrm{T}}_{0}^{{-}1}, \\ \hat{\mathrm{D}}_{1}(t) & =\hat{\boldsymbol{\varTheta}}_{1}(t)\hat{\varGamma}_{1}^{{-}1}\hat{ \boldsymbol{\varXi}}_{1}=\left( \hat{\mathrm{I}},t\hat{\mathrm{I}}\right) \left( \begin{array}{cc} \hat{\mathrm{T}}_{0} & \hat{\mathrm{T}}_{1} \\ \hat{\mathrm{T}}_{1} & \hat{\mathrm{T}}_{2} \end{array} \right)^{{-}1}\left( \begin{array}{c} 0 \\ \hat{\mathrm{I}} \end{array} \right) . \end{aligned} \right\} \end{equation}

Consider the inversion problem

(B7)\begin{equation} \left( \begin{array}{cc} \hat{\mathrm{T}}_{0} & \hat{\mathrm{T}}_{1} \\ \hat{\mathrm{T}}_{1} & \hat{\mathrm{T}}_{2} \end{array} \right) ^{{-}1}\left( \begin{array}{c} 0 \\ \hat{\mathrm{I}} \end{array} \right) =\left( \begin{array}{c} \hat{\mathrm{U}} \\ \hat{\mathrm{V}} \end{array} \right) , \end{equation}

which is equivalent to

(B8)\begin{equation} \left( \begin{array}{cc} \hat{\mathrm{T}}_{0} & \hat{\mathrm{T}}_{1} \\ \hat{\mathrm{T}}_{1} & \hat{\mathrm{T}}_{2} \end{array} \right) \left( \begin{array}{c} \hat{\mathrm{U}} \\ \hat{\mathrm{V}} \end{array} \right) =\left( \begin{array}{c} 0 \\ \hat{\mathrm{I}} \end{array} \right) . \end{equation}

The solution of this system of matrix equations is quite simple,

(B9)\begin{equation} \hat{\mathrm{V}}=\left( \hat{\mathrm{T}}_{2}-\hat{\mathrm{T}}_{1}\mathrm{ \hat{T}}_{0}^{{-}1}\hat{\mathrm{T}}_{1}\right) ^{{-}1}=\hat{\mathrm{C}} ,\quad \hat{\mathrm{U}}={-}\hat{\mathrm{T}}_{0}^{{-}1}\hat{\mathrm{T}}_{1}\hat{\mathrm{V }}={-}\hat{\mathrm{T}}_{0}^{{-}1}\hat{\mathrm{T}}_{1}\hat{\mathrm{C}} , \end{equation}

so that

(B10)\begin{equation} \hat{\mathrm{D}}_{1}(t)=\left( t\hat{\mathrm{I}}-\hat{\mathrm{T}}_{0}^{{-}1} \hat{\mathrm{T}}_{1}\right) \hat{\mathrm{C}} . \end{equation}

The conjugate polynomials are defined as

(B11)\begin{equation} \hat{\mathrm{E}}_{s}\left( z\right) =\frac{1}{{\rm \pi} }\int_{-\infty }^{\infty } \frac{{\rm d}\omega }{\omega -z}\hat{\sigma}_{\mathrm{H}}(\omega )\left( \hat{\mathrm{D}}_{s}(\omega )-\hat{\mathrm{D}}_{s}(z)\right) \end{equation}

and we have that

(B12)\begin{equation} \hat{\mathrm{E}}_{0}(z)=0,\quad \hat{\mathrm{E}}_{1}(z)=\frac{1}{{\rm \pi} } \int_{-\infty }^{\infty }\frac{{\rm d}\omega }{\omega -z}\hat{\sigma}_{\mathrm{H} }(\omega )\left( \hat{\mathrm{D}}_{1}(\omega )-\hat{\mathrm{D}} _{1}(z)\right) =\hat{\mathrm{T}}_{0}\hat{\mathrm{C}}. \end{equation}

The Kramers–Kronig relations (see the Supplemental material) can be recast as

(B13)\begin{equation} \hat{\sigma}(z)=\frac{1}{{\rm \pi} {\rm i}}\int_{-\infty}^{\infty}\hat{\sigma}_{\mathrm{H}}(\omega) \frac{{\rm d}\omega}{\omega - z} , \end{equation}

so that, always for $\mathrm {Im}\ z>0$, and due to the matrix version of the Nevanlinna theorem (Kovalishina Reference Kovalishina1983; Adamyan & Tkachenko Reference Adamyan, Tkachenko and Verlag2000), we obtain the conductivity tensor in a convenient form analogous to that of the scalar case,

(B14)\begin{equation} \hat{\sigma}(z)=\frac{{\rm i}}{{\rm \pi} }\int_{-\infty }^{\infty }\hat{\sigma}_{\mathrm{H}}(\omega )\frac{{\rm d}\omega }{z-\omega }={\rm i}\left( \hat{\mathrm{E}}_{1}(z)+ \hat{\mathrm{E}}_{0}(z)\hat{\mathrm{R}}(z)\right) \left( \hat{\mathrm{D}} _{1}(z)+\hat{\mathrm{D}}_{0}(z)\hat{\mathrm{E}}(z)\right) ^{{-}1}, \end{equation}

where $\hat {\mathrm {R}}(z)$ is the Nevanlinna parameter tensor function $\hat {\mathrm {R}}(z)$ mentioned above. Thus, we arrive at the following solution of the truncated matrix Hamburger problem $\{ \hat {\mathrm {T}}_{0},\hat {\mathrm {T}}_{1},\hat {\mathrm {T}}_{2}\}$ with the third tensor moment $\hat {\mathrm {T}}_{2}$ to be treated as an immaterial element of the solution, i.e. it can be eliminated from the solution by a renormalization of the intrinsically unknown Nevanlinna parameter which possesses specific mathematical properties:

(B15)\begin{align} \hat{\sigma}(z) & ={\rm i}\hat{\mathrm{T}}_{0}\hat{\mathrm{C}}\left[\left(z \hat{\mathrm{I}}-\hat{\mathrm{T}}_{0}^{{-}1}\hat{\mathrm{T}}_{1}\right) \hat{\mathrm{C}}+\hat{\mathrm{T}}_{0}^{{-}1}\hat{\mathrm{R}}(z)\right]^{{-}1} \nonumber\\ & ={\rm i}\hat{\mathrm{T}}_{0}\hat{\mathrm{C}}\left[ \left( z\hat{\mathrm{I}}- \hat{\mathrm{T}}_{0}^{{-}1}\hat{\mathrm{T}}_{1}\right) \hat{\mathrm{C}}+ \hat{\mathrm{T}}_{0}^{{-}1}\hat{\mathrm{R}}(z)\hat{\mathrm{C}}^{{-}1}\hat{\mathrm{C}}\right] ^{{-}1} \nonumber\\ & ={\rm i}\hat{\mathrm{T}}_{0}\hat{\mathrm{C}}\left\{ \left[ \left( z\hat{\mathrm{I }}-\hat{\mathrm{T}}_{0}^{{-}1}\hat{\mathrm{T}}_{1}\right) +\hat{\mathrm{T}} _{0}^{{-}1}\hat{\mathrm{R}}(z)\hat{\mathrm{C}}^{{-}1}\right] \hat{\mathrm{C}} \right\} ^{{-}1}. \end{align}

Then

(B16)\begin{equation} \hat{\sigma}(z)={\rm i}\hat{\mathrm{T}}_{0}\hat{\mathrm{C}}\hat{\mathrm{C}}^{{-}1} \left[ \left( z\hat{\mathrm{I}}-\hat{\varOmega}_{1}+\hat{\mathrm{R}} _{1}(z)\right) \right] ^{{-}1}={\rm i}\hat{\mathrm{T}}_{0}\left[ \left( z\hat{\mathrm{I}}-\hat{\varOmega}_{1}+\hat{\mathrm{R}}_{1}(z)\right) \right] ^{{-}1}, \end{equation}

where

(B17a,b)\begin{equation} \hat{\mathrm{C}}=\left( \hat{\mathrm{T}}_{2}-\hat{\mathrm{T}}_{1}\hat{\mathrm{T}}_{0}^{{-}1}\hat{\mathrm{T}}_{1}\right) ^{{-}1},\quad \hat{\mathrm{R}} _{1}(z)=\hat{\mathrm{T}}_{0}^{{-}1}\hat{\mathrm{R}}(z)\hat{\mathrm{C}}^{{-}1} \end{equation}

and

(B18)\begin{equation} \hat{\varOmega}_{1} = \hat{\mathrm{T}}_{0}^{{-}1}\hat{\mathrm{T}}_{1}. \end{equation}

Consider now the static approximation for the new Nevanlinna parameter

(B19)\begin{equation} \hat{\mathrm{R}}_{1}(z)=\hat{\mathrm{R}}_{1}(0)={\rm i}\hat{\tau}\hat{\varOmega}^{2}, \end{equation}

where $\hat {\tau }$ is the relaxation time tensor. Then the general expression for the conductivity tensor takes the Drude–Lorentz form,

(B20)\begin{equation} \hat{\sigma}(z)=\hat{\mathrm{T}}_{0}\hat{\tau}\left[ \hat{\mathrm{I}} -{\rm i}\left( z\hat{\mathrm{I}}-\hat{\varOmega}_{1}\right) \hat{\tau}\right]^{{-}1}. \end{equation}

The relaxation time tensor can be further determined in terms of the static conductivity. Indeed, since

(B21)\begin{equation} \hat{\sigma}_{0}=\hat{\sigma}(0)=\hat{\mathrm{T}}_{0}\hat{\tau}\left( \hat{\mathrm{I}}+{\rm i}\hat{\varOmega}_{1}\hat{\tau}\right) ^{{-}1} \end{equation}

we have that

(B22)\begin{equation} \hat{\tau}=\left( \hat{\mathrm{T}}_{0}-{\rm i}\hat{\sigma}_{0}\hat{\varOmega} _{1}\right) ^{{-}1}\hat{\sigma}_{0} \end{equation}

and arrive at

(B23)\begin{equation} \hat{\sigma}(z)=\hat{\sigma}_{0}\left( \hat{\mathrm{I}}+{\rm i}\hat{\varOmega}_{1}\hat{\tau}\right) \left[ \hat{\mathrm{I}}-{\rm i}\left( z\hat{\mathrm{I}}- \hat{\varOmega}_{1}\right) \hat{\tau}\right] ^{{-}1}. \end{equation}

The dynamic conductivity tensor (B23) is expressed in terms of the static conductivity tensor (5.31) or (5.32) (see Rylyuk & Tkachenko) and the relaxation time tensor $\hat {\tau }=\hat {\nu }^{-1}$ ($\hat {\nu }$ is the collision frequency tensor). These results lead to the Drude–Lorentz-type formula for the conductivity tensor of a magnetized plasma. Further generalizations are relatively cumbersome and they will be considered elsewhere. The same should be said with respect to the possibilities to avoid the static approximation of the Nevanlinna parameter function (B19).

Appendix C. Static electrical conductivity tensor of a 3-D Dirac plasma in a magnetic field

The Cartesian components of the static electrical conductivity tensor (5.29) of a 3-D Dirac plasma in a magnetic field are

(C1)\begin{gather} \left. \begin{aligned} \sigma_{zz,+} & = \sigma_{zz,1,+} + \sigma_{zz,2,+}, \ \sigma_{zz,-} = \sigma_{zz, 1,-} , \\ \sigma_{xx, \pm} & ={-}\omega_{\rm H}^2\frac{\hbar^3 v^2}{R_k}\int_{-\infty}^{\infty}\sum_{n = 0}^{\infty}\left[\mathfrak{F}(\epsilon_{n,\pm}(k_z)) + \mathfrak{F}(\epsilon_{n+1,\pm}(k_z))\right] \mathcal{J}_{n,n+1,\pm}(x,x,k_z)\,{\rm d}k_z , \\ \sigma_{yy, \pm} & ={-}\omega_{\rm H}^2\frac{\hbar^3 v^2}{R_k}\int_{-\infty}^{\infty}\sum_{n = 0}^{\infty}\left[\mathfrak{F}(\epsilon_{n,\pm}(k_z)) + \mathfrak{F}(\epsilon_{n+1,\pm}(k_z))\right] \mathcal{J}_{n,n+1,\pm}(y,y,k_z)\,{\rm d}k_z , \\ \sigma_{xy, \pm} & ={-}\omega_{\rm H}^2\frac{\hbar^3 v^2}{R_k}\int_{-\infty}^{\infty}\sum_{n = 0}^{\infty}\left[\mathfrak{F}(\epsilon_{n,\pm}(k_z)) - \mathfrak{F}(\epsilon_{n+1,\pm}(k_z))\right] \mathcal{J}_{n,n+1,\pm}(x,y,k_z)\,{\rm d}k_z , \\ \sigma_{zz,2,+} & ={-}2\omega_{\rm H}^4\frac{\hbar^5 v^2}{R_k}\int_{-\infty}^{\infty}\sum_{n = 0}^{\infty}\left[\mathfrak{F}(\epsilon_{n,+}(k_z)) + \mathfrak{F}(\epsilon_{n+2,+}(k_z))\right] \mathcal{J}_{n,n+2,\pm}(z,z,k_z)\,{\rm d}k_z ,\\ \sigma_{zz,1, \pm} & ={-}\frac{\hbar \omega_{\rm H}^2}{2{\rm \pi} R_k \nu_{{\pm}}}\int_{-\infty}^{\infty} \sum_{n=0}^{\infty}\mathfrak{F}(\epsilon_{n,\pm}(k_z))J_{z,n,0,\pm}^2(k_z)\,{\rm d}k_z , \end{aligned} \right\} \end{gather}
(C2a,b)\begin{gather} \mathfrak{F}(\epsilon_{n,\pm}(k_z)) = \left. \frac{\partial f_{\rm FD}(\epsilon)} {\partial \epsilon}\right|_{\epsilon = \epsilon_{n,\pm}(k_z)} , \quad \mathcal{J}_{n,m,\pm}(i,j,k_z) = \frac{\nu_{{\pm}} J_{i,n,\pm}(k_z) J_{j,n,\pm}(k_z)} {\omega_{n,m,\pm}^2(k_z) + \nu_{{\pm}}^2} , \end{gather}

where $\omega _{n,m,\pm }(k_z) = [\epsilon _{n,\pm }(k_z) - \epsilon _{m,\pm }(k_z)]/\hbar$, $R_k = 2{\rm \pi} \hbar /e^2 = 25812.80745\ldots$ $\varOmega$ is the von Klitzing constant or the quantum of electrical resistance and $J_{i,n,\pm }(k_z)$ ($i = x,y,z$) are given by (A2).

Appendix D. Normalization in a magnetic field

Here we calculate the chemical potential of a 3-D and 2-D Fermi fluid of a Dirac one-component plasma.

Since the 3-D number of states in a magnetic field is

(D1)\begin{equation} \frac{eHV}{4{\rm \pi}^2\hbar^2 c} , \end{equation}

the normalization condition in a magnetic field reads

(D2)\begin{equation} \frac{eHV}{4{\rm \pi}^2 \hbar c}\sum_{n=0}^{\infty}\int_{-\infty}^{\infty} \left\{\frac{1}{\exp(\beta \epsilon_{n,+}(k_z) - \eta) + 1} + \frac{1}{\exp(\beta \epsilon_{n,-}(k_z) - \eta) + 1}\right\}\,{\rm d}k_{z} = N_3 , \end{equation}

where $N_3$ is the total number of particles and $\epsilon _{n,+}$ and $\epsilon _{n,-}$ are given by (3.6a,b). This normalization condition can also be represented in the form

(D3)\begin{equation} \frac{1}{2{\rm \pi}^2 a_{\rm H}^2}\left\{-\frac{1}{l}\int_0^{\infty} \frac{{\rm d}t}{\exp(t - \eta) + 1} + \sum_{n=0}^{\infty}\int_{-\infty}^{\infty} \frac{{\rm d}k_{z}}{\exp(\beta \epsilon_{n,-}(k_z) - \eta) + 1}\right\} = n_3 , \end{equation}

where $a_\textrm {H} = \sqrt {c\hbar /(eH)}$ being the magnetic length, $l = \beta hv$ and $n_3 = N_3/V$. In the limit $H \rightarrow 0$ we may replace the summation in (D3) by integration using the Euler–Maclaurin sum formula

(D4)\begin{equation} \sum_{a \leq n < b}F(n)\simeq\int_a^b F(x)\,{{\rm d} x}. \end{equation}

Then $\hbar ^2 \omega _\textrm {H}^2 n\rightarrow \hbar ^2 \omega _\textrm {H}^2 x\rightarrow v^2 h^2 k_{\perp }^2$ and ${\textrm {d} x} = 2v^2/\omega _\textrm {H}^2 k_{\perp }\,\textrm {d}k_{\perp }$. Taking into account that in the limit $H \rightarrow 0$, $\epsilon _{n,-}(k_z) = \hbar \sqrt {\omega _\textrm {H}^2 n + v^2 k_z^2} \rightarrow \hbar v \sqrt {k_{\perp }^2 + k_z^2} = \hbar vk$, we obtain

(D5)\begin{equation} \frac{eH}{{\rm \pi}^2\hbar c}\frac{v^2}{\omega_{\rm H}^2}\int_{0}^{\infty}k_{{\perp}}\,{\rm d}k_{{\perp}} \int_{-\infty}^{\infty}\frac{{\rm d}k_z}{\exp(\beta vhk - \eta) + 1} = n_3. \end{equation}

But $2{\rm \pi} k_{\perp }\,\textrm {d}k_{\perp }\,\textrm {d}k_z = 4{\rm \pi} k^2 \,\textrm {d}k$ and then

(D6)\begin{equation} \frac{1}{{\rm \pi}^2}\int_0^{\infty}\frac{k^2 \,{\rm d}k}{\exp(\beta vhk - \eta) + 1} = n_3. \end{equation}

Making a change of variables $\beta vhk = t$ we arrive at the result

(D7)\begin{equation} \int_0^{\infty}\frac{t^2 \,{\rm d}t}{\exp(t - \eta) + 1} = {\rm \pi}^2 l^3 n_3. \end{equation}

So, the dimensionless chemical potential $\eta$ for 3-D massless Dirac fermions in the absence of a magnetic field is determined by the condition (D7). In non-degenerate, hot plasmas ($\beta \rightarrow 0$), the Fermi–Dirac distribution in (D7) can be substituted by the Maxwell one and we have

(D8)\begin{equation} \int_0^{\infty}\frac{t^2 \,{\rm d}t}{\exp(t - \eta) + 1} \simeq {\rm e}^{\eta} \int_0^{\infty}t^2 {\rm e}^{{-}t}\,{\rm d}t = 2{\rm e}^{\eta} \simeq {\rm \pi}^2 l^3 n_3. \end{equation}

Then for non-degenerate hot 3-D Dirac plasmas, in the absence of a magnetic field, $\eta \simeq \ln ({\rm \pi} ^2 l^3 n_3/2)$.

We can also get an analytical approximation for the dimensionless chemical potential $\eta$ for non-degenerate hot 3-D Dirac plasmas, in the case of strong magnetic fields. Under the condition $1\lesssim \beta \hbar \omega _\textrm {H} \ll \eta$, from (D3), we obtain

(D9)\begin{equation} \frac{{\rm e}^{\eta}}{2{\rm \pi}^2 a_{\rm H}^2}\left\{\frac{1}{l} + 2\frac{\omega_{\rm H}}{v} \sum_{n=1}^{\infty}\sqrt{n}\ {\rm K}_1(\beta \hbar \omega_{\rm H} \sqrt{n})\right\}\simeq n_3 \end{equation}

and then

(D10)\begin{equation} \eta \simeq \ln \left\{\frac{2{\rm \pi}^2 a_{\rm H}^2 n_3}{1/l + 2(\omega_{\rm H}/v) \sum_{n=1}^{\infty}\sqrt{n}\ {\rm K}_1(\beta \hbar \omega_{\rm H}\sqrt{n})}\right\} , \end{equation}

where

(D11)\begin{equation} {\rm K}_{\nu}(z) = \int_0^{\infty}{\rm e}^{{-}z\coth(t)}\coth(\nu t)\,{\rm d}t \end{equation}

is the Macdonald function. Let us give some numerical estimates. For $T = 10^4$ K, $n_3 = 10^{19}$ cm$^{-3}$, $H = 30$ MG, $v \simeq c/300$; $\beta \hbar \omega _\textrm {H} \simeq 2.3$ and $\eta \simeq -4$. For $T = 10^5$ K, $n_3 = 10^{21}$ cm$^{-3}$, $H = 700$ MG, $v \simeq c/300$; $\beta \hbar \omega _\textrm {H} \simeq 1.11$ and $\eta \simeq -6.14$.

Taking into account (3.9) the normalization condition for 2-D massless Dirac fermions in a magnetic field is

(D12)\begin{equation} \frac{eHS}{2{\rm \pi} \hbar c}\sum_{n=0}^{\infty} \left\{\frac{1}{\exp(\beta\epsilon_{n,+} - \eta) + 1} + \frac{1} {\exp(\beta\epsilon_{n,-} - \eta) + 1}\right\} = N_2 , \end{equation}

where $\epsilon _{n,+}$ and $\epsilon _{n,-}$ are given by (3.8a,b). This normalization condition can also be represented in the form

(D13)\begin{equation} \frac{1}{2{\rm \pi} a_{\rm H}^2}\left\{-\frac{1}{\exp(- \eta) + 1} + 2\sum_{n=0}^{\infty} \frac{1}{\exp(\beta\epsilon_{n,-} - \eta) + 1}\right\} = n_2 , \end{equation}

where $n_2 = N_2/S$. In the limit $H \rightarrow 0$, in view of (D4), from (D13), we obtain

(D14)\begin{equation} \frac{1}{\rm \pi}\int_0^{\infty}\frac{k_{{\perp}}\,{\rm d}k_{{\perp}}}{\exp(\beta vhk_{{\perp}} - \eta) + 1} = n_2 . \end{equation}

Making a change of variables $\beta v h k_{\perp } = t$ we have

(D15)\begin{equation} \int_0^{\infty}\frac{t\,{\rm d}t}{\exp(t - \eta) + 1} = {\rm \pi}l^2 n_2 . \end{equation}

The dimensionless chemical potential $\eta$ for 2-D massless Dirac fermions in the absence of a magnetic field is determined by (D15). In non-degenerate, hot plasmas ($\beta \rightarrow 0$), the Fermi–Dirac distribution in (D15) can be substituted by the Maxwell one and we have

(D16)\begin{equation} \int_0^{\infty}\frac{t\,{\rm d}t}{\exp(t - \eta) + 1} \simeq {\rm e}^{\eta} \int_0^{\infty}t{\rm e}^{{-}t}\,{\rm d}t = {\rm e}^{\eta} \simeq {\rm \pi}l^2 n_2 . \end{equation}

Then for non-degenerate hot 2-D Dirac plasmas, in the absence of a magnetic field, $\eta \simeq \ln ({\rm \pi} l^2 n_2)$.

We can also get an analytical approximation for the dimensionless chemical potential $\eta$ for non-degenerate hot 2-D Dirac plasmas, in the case of strong magnetic fields. Under the condition $1\lesssim \beta \hbar \omega _\textrm {H} \ll \eta$, from (D13), we obtain

(D17)\begin{equation} \frac{{\rm e}^{\eta}}{2{\rm \pi} a_{\rm H}^2}\left\{1 + 2\sum_{n=1}^{\infty}\exp(-\beta\epsilon_{n,-}) \right\}\simeq n_2 \end{equation}

and then

(D18)\begin{equation} \eta \simeq \ln\left\{\frac{2{\rm \pi} a_{\rm H}^2 n_2}{1 + 2\sum_{n=1}^{\infty} \exp(-\beta \hbar \omega_{\rm H} \sqrt{n})}\right\} . \end{equation}

Let us give some numerical estimates. For $T = 10^5$ K, $n_2 = 10^{14}$ cm$^{-2}$, $H = 700$ MG, $v \simeq c/300$; $\beta \hbar \omega _\textrm {H} \simeq 1.11$ and $\eta \simeq -4.1$. For $T = 10^5$ K, $n_2 = 10^{15}$ cm$^{-2}$, $H = 700$ MG, $v \simeq c/300$; $\beta \hbar \omega _\textrm {H} \simeq 1.11$ and $\eta \simeq -1.8$.

References

Abrikosov, A.A. 1988 Fundamentals of the Theory of Metals. Elsevier Science Publishers B. V.Google Scholar
Adamyan, V.M. & Tkachenko, I.M. 2000 Solution of the truncated matrix Hamburger moment problem according to M.G. Krein. In Operator Theory: Advances and Applications (ed. Verlag, Birkhäuser), vol. 118, pp. 3351. Birkhäuser.Google Scholar
Akhiezer, N.I. 1965 The Classical Moment Problem and Some Related Questions in Analysis. Hafner Publishing Company.Google Scholar
Akhiezer, A.I. & Peletminskii, S.V. 1981 Methods of Statistical Physics. Pergamon Press.Google Scholar
Andersen, K., Jacobsen, K.W. & Thygesen, K.S. 2014 Plasmons on the edge of $\textrm {MoS}_2$ nanostructures. Phys. Rev. B 90, 161410(R).CrossRefGoogle Scholar
Berestetskii, V.B., Pitaevskii, L.P. & Lifshitz, E.M. 1982 Quantum Electrodynamics, 2nd edn. Course of Theoretical Physics, vol. 4. Butterworth-Heinemann.Google Scholar
Bjorken, J.D. & Drell, S.D. 1964 Relativistic Quantum Mechanics. McGraw-Hill.Google Scholar
Blank, A.Y. & Kaner, E.A. 1966 The phonon spectrum of metals in a magnetic field. Zh. Eksp. Teor. Fiz. 50, 1013 [Sov. Phys. JETP 23, 673 (1966)].Google Scholar
Castro Neto, A.H., Guinea, F., Peres, N.M.R., Novoselov, K.S. & Geim, A.K. 2009 The electronic properties of graphene. Rev. Mod. Phys. 81, 109162.CrossRefGoogle Scholar
Enaldiev, V.V. 2018 Collective excitations in a two-component one-dimensional massless Dirac plasma. Phys. Rev. B 98, 155417-1155417-8.CrossRefGoogle Scholar
Heinz, U. 2009 The strongly coupled quark–gluon plasma created at RHIC. J. Phys. A 42, 214003214010; Thoma, M. H. 2009 What can we learn from electromagnetic plasmas about the quark–gluon plasma? J. Phys. A 42, 214004–214015.CrossRefGoogle Scholar
Katsnelson, M.I., Novoselov, K.S. & Geim, A.K. 2006 Chiral tunnelling and the Klein paradox in graphene. Nat. Phys. 2 (9), 620625.CrossRefGoogle Scholar
Kovalishina, I.V. 1983 New aspects of the classical problem of moments. Izv. Akad. Nauk. Ser. Mat. 47, 455 [Math. USSR Izv. 22, 419 (1984)].Google Scholar
Krein, M.G. & Nudel'man, A.A. 1977 The Markov Moment Problem and Extremal Problems. Trans. of Math. Monographs 50. American Mathematical Society.CrossRefGoogle Scholar
Landau, L.D. & Lifshitz, E.M. 1980 Statistical Physics, Pt.1. Course of Theoretical Physics, vol. 5. Pergamon Press.Google Scholar
McCann, E. & Fal'ko, V.I. 2006 Landau-level degeneracy and quantum hall effect in a graphite bilayer. Phys. Rev. Lett. 96, 086805.CrossRefGoogle Scholar
Moldabekov, Z., Ludwig, P., Bonitz, M. & Ramazanov, T. 2015 Ion potential in warm dense matter: wake effects due to streaming degeneratee lectrons. Phys. Rev. E 91, 023102-1023102-16 and references therein.CrossRefGoogle Scholar
Novoselov, K.S., Geim, A.K., Morozov, S.V., Jiang, D., Katsnelson, M.I., Grigorieva, I.V., Dubonos, S.V. & Firsov, A.A. 2005 Two dimensional gas of massless Dirac fermions in graphene. Nature 438, 197.CrossRefGoogle ScholarPubMed
Ren, Y.-N., Cheng, Q., Li, S.-Y., Chao, Y., Liu, Y.-W., Lv, K., Zhang, M.-H., Sun, Q.-F. & He, L. 2021 Spatial and magnetic confiement of massless Dirac fermions. Phys. Rev. B 104, L161408-1L161408-6.CrossRefGoogle Scholar
Sarma, S.D. & Hwang, E.H. 2009 Collective modes of the massless Dirac plasma. Phys. Rev. Lett. 102, 206412-1206412-4.Google ScholarPubMed
Sarri, G., Macchi, A., Cecchetti, C.A., Kar, S., Liseykina, T.V., Yang, X.H., Dieckmann, M.E., Fuchs, J., Galimberti, M., Gizzi, L.A, et al. 2012 Dynamics of self-generated, large amplitude magnetic fields following high-intensity laser matter interaction. Phys. Rev. Lett. 109, 205002-1205002-5.CrossRefGoogle ScholarPubMed
Shaisultanov, R., Lyubarsky, Y. & Eichler, D. 2012 Stream instabilities in relativistically hot plasma. Astrophys. J. 744, 182189.CrossRefGoogle Scholar
Skobov, V.G. & Kaner, E.A. 1964 Quantum theory of the propagation of electromagnetic waves in metals in a magnetic gield. Zh. Eksp. Teor. Fiz. 46, 1809 [Sov. Phys. JETP 19, 1219 (1964)].Google Scholar
Tatarakis, M., Watts, I., Beg, F.N., Clark, E.L., Dangor, A.E., Gopal, A., Haines, M.G., Norreys, P.A., Wagner, U., Wei, M.-S., Zepf, M. & Krushelnick, K. 2002 a Measuring huge magnetic fields. Nature 415, 280.CrossRefGoogle ScholarPubMed
Tatarakis, M., Gopal, A., Watts, I., Beg, F.N., Wei, M.S., Dangor, A.E., Krushelnick, K., Wagner, U., Norreys, P.A., Clark, E.L., Zepf, M. & Evans, R.G. 2002 b Measurements of ultrastrong magnetic fields during relativistic laser–plasma interactions. Phys. Plasmas 9, 36423642.CrossRefGoogle Scholar
Wagner, U., Tatarakis, M., Gopal, A., Beg, F.N., Clark, E.L., Dangor, A.E., Evans, R.G., Haines, M.G., Mangles, S.P.D., Norreys, P.A., et al. 2004 Laboratory measurements of 0.7 GG magnetic fields generated during high-intensity laser interactions with dense plasmas. Phys. Rev. E 70, 026401026406.CrossRefGoogle ScholarPubMed
Yamamoto, N. & Yang, D.-L. 2021 Helical magnetic effect and the chiral anomaly. Phys. Rev. D 103, 125003-1125003-7.CrossRefGoogle Scholar
Zubarev, D.N. 1974 Nonequilibrium Statistical Thermodynamic. Plenum.Google Scholar
Figure 0

Figure 1. The function $\mathfrak {M}^{*}$ versus $H$ for $n_3 = 10^{22}$ cm$^{-3}$, $T = 10$ K and $v \approx c/300$.

Figure 1

Figure 2. The function $\mathfrak {M}_\textrm {2D}$ versus $H$ for $n_2 = 10^{18}$ cm$^{-2}$, $T = 100$ K and $v \approx c/300$.

Figure 2

Figure 3. The function $\sigma$ versus $H$ for $n_3 = 10^{21}$ cm$^{-3}$, $T = 10$ K and $v \approx c/300$.

Figure 3

Figure 4. The function $R_{zz}$ versus $H$ for $n_3 = 10^{21}$ cm$^{-3}$, $T = 100$ K and $v \approx c/300$.

Figure 4

Figure 5. The function $\sigma _\textrm {2D}$ versus $H$ for $n_2 = 10^{19}$ cm$^{-2}$, $T = 10$ K and $v \approx c/300$.

Supplementary material: PDF

Rylyuk and Tkachenko supplementary material

Rylyuk and Tkachenko supplementary material 1

Download Rylyuk and Tkachenko supplementary material(PDF)
PDF 66.3 KB
Supplementary material: File

Rylyuk and Tkachenko supplementary material

Rylyuk and Tkachenko supplementary material 2

Download Rylyuk and Tkachenko supplementary material(File)
File 115.3 KB