Hostname: page-component-848d4c4894-m9kch Total loading time: 0 Render date: 2024-05-01T16:47:13.330Z Has data issue: false hasContentIssue false

Hall magnetohydrodynamics in a relativistically strong mean magnetic field

Published online by Cambridge University Press:  12 December 2023

Y. Kawazura*
Affiliation:
Frontier Research Institute for Interdisciplinary Sciences, Tohoku University, 6-3 Aoba, Aramaki, Sendai 980-8578, Japan Department of Geophysics, Graduate School of Science, Tohoku University, 6-3 Aoba, Aramaki, Aoba-ku, Sendai 980-8578, Japan
*
Email address for correspondence: kawazura@tohoku.ac.jp
Rights & Permissions [Opens in a new window]

Abstract

This paper presents a magnetohydrodynamic model that describes the small-amplitude fluctuations with wavelengths comparable to ion inertial length in the presence of a relativistically strong mean magnetic field. The set of derived equations is virtually identical to the non-relativistic Hall reduced magnetohydrodynamics (Schekochihin et al., J. Plasma Phys., vol. 85, 2019, 905850303), differing only by a few constants that take into account the relativistic corrections. This means that all the properties of kinetic Alfvén turbulence and ion cyclotron turbulence inherent in the non-relativistic Hall regime persist unchanged even in a magnetically dominated regime.

Type
Letter
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press

1. Introduction

Turbulence of relativistically magnetized plasmas (here defined as the magnetic field energy exceeding the rest mass energy of particles) can be found in a number of astrophysical systems, e.g. pulsar and black hole magnetospheres, coronae of accretion disks, and jets from active galactic nuclei. The turbulent fluctuations of the magnetic field in these systems can be dissipated and converted into the thermal and non-thermal energy of particles (Zhdankin et al. Reference Zhdankin, Walker, Boldyrev and Lesur2017; Comisso & Sironi Reference Comisso and Sironi2018; Zhdankin et al. Reference Zhdankin, Uzdensky, Werner and Begelman2019; Nättilä & Beloborodov Reference Nättilä and Beloborodov2022), which are potential sources of the bright electromagnetic radiation we observe on the Earth. As the reservoir of magnetic energy is huge, even small-amplitude fluctuations give rise to significant heating and acceleration. Thus, understanding the properties of turbulent fluctuations in relativistically magnetized plasmas (also known as magnetically dominated plasmas) is one of the most important themes in modern high-energy astrophysics.

In the vast majority of studies of relativistic turbulence, either ideal magnetohydro-dynamics (MHD) or fully kinetic Vlasov–Maxwell equations are used (with a few exceptions that used the resistive MHD to describe proper reconnection (e.g. Ripperda, Bacchini & Philippov Reference Ripperda, Bacchini and Philippov2020; Ripperda et al. Reference Ripperda, Liska, Chatterjee, Musoke, Philippov, Markoff, Tchekhovskoy and Younsi2022) and that used relativistic Braginskii equations to include weakly collisional effects (e.g. Foucart et al. Reference Foucart, Chandra, Gammie and Quataert2016, Reference Foucart, Chandra, Gammie, Quataert and Tchekhovskoy2017)). Ideal MHD is, by definition, only able to describe the large-scale dynamics, and thus, it is not suitable for studying the dissipation of fluctuations, which usually occurs on scales smaller than the ion inertial length or the ion Larmor radius. Vlasov–Maxwell equations, on the other hand, can properly describe the dissipation processes, but it is a rather too complex model. In fact, when Vlasov–Maxwell equations are used for relativistic turbulence, they are solved only by means of ab initio particle-in-cell (PIC) simulations, while the analytical study of small-scale physics of relativistic plasmas is underdeveloped. The aim of this study is to formulate a comprehensive and useful relativistic magnetohydrodynamic model that is valid even at small scales.

The extension of relativistic ideal MHD to incorporate the small-scale effects was first proposed by Koide (Reference Koide2009) (which was then rederived using a variational principle (Kawazura, Miloshevich & Morrison Reference Kawazura, Miloshevich and Morrison2017)). The model is often called relativistic extended MHD (XMHD), which includes the Hall effect, the rest mass inertia of electrons and the thermal inertia of electrons. When the inertia effects of electrons are neglected, this set of equations is referred to as relativistic Hall MHD (HMHD). To date, relativistic XMHD and relativistic HMHD have been widely used, e.g. for studying magnetic reconnection (Comisso & Asenjo Reference Comisso and Asenjo2014, Reference Comisso and Asenjo2018; Asenjo & Comisso Reference Asenjo and Comisso2019; Yang Reference Yang2019b,Reference Yanga), magnetofluid topological connection (Asenjo & Comisso Reference Asenjo and Comisso2015; Asenjo, Comisso & Mahajan Reference Asenjo, Comisso and Mahajan2015; Comisso & Asenjo Reference Comisso and Asenjo2020) and linear wave propagation (Kawazura Reference Kawazura2017, Reference Kawazura2022). However, these models have not been used for turbulence.Footnote 1 Since the set of relativistic XMHD equations is much more complicated than that of non-relativistic XMHD or relativistic ideal MHD, it may be too difficult to solve relativistic XMHD as it is, even using direct numerical simulations. Alternatively, in this work, we reduce relativistic HMHD to make it more tractable by assuming the presence of a mean magnetic field – a technique commonly used for non-relativistic models.

When the spatial scale of the turbulent fluctuations is much smaller than the scale of energy injection (which is macroscopic in many astrophysical systems), the large-scale magnetic field effectively behaves like a mean field for the fluctuations (Kraichnan Reference Kraichnan1965; Howes et al. Reference Howes, Cowley, Dorland, Hammett, Quataert and Schekochihin2008). Therefore, as the turbulent cascade progresses, the fluctuations become smaller amplitude and more elongated along the mean field. Consequently, the ideal MHD asymptotically becomes reduced MHD (RMHD) (Kadomtsev & Pogutse Reference Kadomtsev and Pogutse1974; Strauss Reference Strauss1976). While non-relativistic RMHD has been widely used in studies of magnetically confined fusion, solar wind (e.g. Chen et al. Reference Chen, Mallet, Yousef, Schekochihin and Horbury2011), planetary magnetospheres (e.g. Watanabe Reference Watanabe2010) and accretion flows (Kawazura et al. Reference Kawazura, Schekochihin, Barnes, Dorland and Balbus2022), relativistic RMHD was formulated only recently (Chandran, Foucart & Tchekhovskoy Reference Chandran, Foucart and Tchekhovskoy2018; TenBarge et al. Reference TenBarge, Ripperda, Chernoglazov, Bhattacharjee, Mahlmann, Most, Juno, Yuan and Philippov2021). Remarkably, relativistic RMHD and non-relativistic RMHD are formally identical except for the definition of Alfvén speed which is modified such that it never exceeds the speed of light. This means that all the properties of turbulence described by non-relativistic RMHD are true even in the relativistic regime (for example, the Alfvén and slow waves are energetically decoupled (Schekochihin et al. Reference Schekochihin, Cowley, Dorland, Hammett, Howes, Quataert and Tatsuno2009) while the fast waves are entirely ordered out; see TenBarge et al. (Reference TenBarge, Ripperda, Chernoglazov, Bhattacharjee, Mahlmann, Most, Juno, Yuan and Philippov2021) for a detailed discussion on relativistic RMHD).

In the non-relativistic regime, the same reduction procedure can be adopted for Hall MHD, and the resulting model is called Hall reduced MHD (HRMHD) (Gómez, Mahajan & Dmitruk Reference Gómez, Mahajan and Dmitruk2008), which is valid at the ion inertial length. The HRMHD can also be derived by gyrokinetics when ions are cold, and the electron beta is order unity (Schekochihin, Kawazura & Barnes (Reference Schekochihin, Kawazura and Barnes2019), hereafter S19). The Alfvén waves and slow waves (which are decoupled in RMHD) are reorganized into the kinetic Alfvén waves (KAW) and oblique ion cyclotron waves (ICW) in HRMHD.

Now, it is quite natural to ask whether the relativistic effects alter the properties of KAW and ICW turbulence in the HRMHD limit. Here, we formulate the relativistic version of HRMHD (i.e. the relativistic extension of S19, or equivalently, the inclusion of the Hall effect in Chandran et al. (Reference Chandran, Foucart and Tchekhovskoy2018) and TenBarge et al. (Reference TenBarge, Ripperda, Chernoglazov, Bhattacharjee, Mahlmann, Most, Juno, Yuan and Philippov2021)). This is a simple and comprehensive model that is valid at the microscopic scales when the background magnetic field is relativistically strong. We find that the relativistic HRMHD is almost identical to the non-relativistic HRMHD (S19), and thus the properties of KAW and ICW in a non-relativistic regime are also true in a magnetically dominated regime.

2. Derivation of relativistic HRMHD

Consider quasineutral relativistic ion and electron fluids with infinitely small electron-to-ion mass ratio. To describe the time evolution of such plasmas, we use relativistic HMHD (Kawazura Reference Kawazura2017), which consists of the mass conservation law

(2.1a)\begin{equation} \frac{\partial }{\partial t}\left( n\gamma \right) + \boldsymbol{\nabla}\boldsymbol{\cdot}\left( n\gamma\boldsymbol{u} \right) = 0, \end{equation}

the momentum equation

(2.1b)\begin{equation} \frac{\partial }{\partial t}\left( nh\gamma^2\boldsymbol{u} \right) + \boldsymbol{\nabla}\boldsymbol{\cdot}\left( nh\gamma^2\boldsymbol{u}\boldsymbol{u} \right) ={-}c^2\boldsymbol{\nabla} p + c^2\rho_\mathrm{q}\boldsymbol{E} + c\boldsymbol{J}\times\boldsymbol{B}, \end{equation}

the generalized Ohm's law

(2.1c)\begin{equation} \boldsymbol{E} + \frac{\boldsymbol{u}}{c}\times\boldsymbol{B} = \frac{1}{\gamma e n}\left( \rho_\mathrm{q} \boldsymbol{E} + \frac{\boldsymbol{J}}{c}\times\boldsymbol{B} - \boldsymbol{\nabla} p_\mathrm{e} \right) \end{equation}

and Maxwell's equations

(2.1d)\begin{gather} \boldsymbol{\nabla}\boldsymbol{\cdot}\boldsymbol{E} = 4{\rm \pi}\rho_\mathrm{q}, \end{gather}
(2.1e)\begin{gather}-\frac{\partial \boldsymbol{E}}{\partial t} + c\boldsymbol{\nabla}\times\boldsymbol{B} = 4{\rm \pi}\boldsymbol{J}, \end{gather}
(2.1f)\begin{gather}\boldsymbol{\nabla}\boldsymbol{\cdot}\boldsymbol{B} = 0, \end{gather}
(2.1g)\begin{gather}\frac{\partial \boldsymbol{B}}{\partial t} + c\boldsymbol{\nabla}\times\boldsymbol{E} = 0, \end{gather}

where $e$ is the elementary charge, $c$ is the speed of light, $n$ is the rest frame number density, $h$ is the total thermal enthalpy, $p$ is the total thermal pressure, $p_\mathrm{e}$ is the thermal pressure of electrons, $\boldsymbol {u}$ is the flow velocity, $\gamma = 1/\sqrt {1 - |\boldsymbol {u}|^2/c^2}$ is the Lorentz factor, $\boldsymbol {J}$ is the electric current, $\rho _\mathrm{q}$ is the charge density, $\boldsymbol {E}$ is the electric field and $\boldsymbol {B}$ is the magnetic field. The relativistic ideal MHD is recovered when the right-hand side of (2.1c) is neglected.

In what follows, we assume that all fields are separable into spatiotemporally constant background (symbols with a subscript 0) and fluctuations (symbols with $\delta$ in front), viz. $n = n_0 + \delta n, \boldsymbol {B} = (B_0 + \delta B_{\|})\hat {\boldsymbol {z}} + \delta \boldsymbol {B}_{\bot }$, and so on. Electrons are assumed to be isothermal, i.e. $\delta p_{{\rm e}} = T_{{{\rm e}}0}\delta n$, where $T_{{{\rm e}}0}$ is the background electron temperature. Here, $\hat {\boldsymbol {z}} = \boldsymbol {B}_0/|\boldsymbol {B}_0|$, and $\| (\bot )$ denotes the parallel (perpendicular) component to $\boldsymbol {B}_0$. We also assume that the mean flow is absent, i.e. $\boldsymbol {u}_0 = 0$. Plugging the constant background fields into (2.1c), (2.1d) and (2.1e), one finds $\boldsymbol {E}_0 = \boldsymbol {J}_0 = \rho _{\mathrm{q}0} = 0$. Then, we impose the reduced MHD ordering

(2.2)\begin{equation} \frac{\delta n}{n_0} \sim \frac{\delta \boldsymbol{B}}{B_0} \sim \frac{\boldsymbol{u}}{v_\mathrm{A}} \sim \frac{\delta p_\mathrm{e}}{p_{\mathrm{e}0}} \sim \frac{k_{\|}}{k_{\bot}} \sim \epsilon \ll 1, \quad \frac{\partial }{\partial t} \sim \omega \sim k_{\|} v_\mathrm{A}, \end{equation}

where $\omega$ and $\boldsymbol {k}$ are the frequency and wavenumber of the fluctuations, respectively. Here, we have defined the relativistic Alfvén speed,

(2.3)\begin{equation} v_\mathrm{A} = \frac{cB_0}{\sqrt{4{\rm \pi} n_0 h_0 + B_0^2}} = \sqrt{\frac{\sigma}{1 + \sigma}}\,c, \end{equation}

where $\sigma = B_0^2/4{\rm \pi} n_0 h_0$ is the magnetization parameter. When $\sigma \gtrsim 1$ (equivalently $v_\mathrm{A} \approx c$), the plasma is relativistically magnetized. Since $\boldsymbol {u}$ is small, the relativistic effect of bulk flow is absent, i.e. $\gamma \approx 1$, but this is acceptable because we are interested in the microscopic scales where bulk flow is generally small while the thermal energy and/or magnetic energy can be relativistic. We also assume that the ions are cold while the electron can be relativistically hot,

(2.4)\begin{equation} T_{\mathrm{i}0} \ll T_{\mathrm{e}0} \ll \sqrt{\frac{m_\mathrm{i}}{m_\mathrm{e}}}\,m_\mathrm{e} c^2 \approx 20\,\mathrm{MeV}, \end{equation}

where $T_{s0}$ and $m_s$ are temperature and mass of the species $s$. The upper bound for the electron temperature is required so that the relativistic thermal inertia of electrons is negligible at the ion inertial scale. Since the ions are cold, the thermal inertia of the ions is also negligible, i.e. $h_0 \approx m_\mathrm{i} c^2$. These conditions enforce the restriction,

(2.5)\begin{equation} \beta_\mathrm{e} = \left(\frac{2T_{\mathrm{e}0}}{m_\mathrm{i} c^2}\right)\frac{1}{\sigma} \ll \sqrt{\frac{m_\mathrm{e}}{m_\mathrm{i}}}, \end{equation}

where $\beta _\mathrm{e} = 8{\rm \pi} p_{\mathrm{e}0}/B_0^2$ is the electron beta. This is different from the assumptions $T_{\mathrm{i}0} \ll T_{\mathrm{e}0}$ and $\beta _\mathrm{e} \sim 1$ that are used in the non-relativistic HRMHD (S19) because they cannot be satisfied when the magnetization is relativistic, i.e. $\sigma \gtrsim 1$. In this work, we consider the electron beta to be lower than that of non-relativistic HMHD, which is formally allowed as long as $\epsilon \ll \beta _\mathrm{e}$. In other words, we treat $\beta _\mathrm{e}$ as order unity, although it is much smaller than $\sqrt {m_\mathrm{e}/m_\mathrm{i}}$, because $\epsilon$ is assumed to be even smaller than $\beta _\mathrm{e}$. However, $\beta _\mathrm{e}$ may not be too small because when $\beta _\mathrm{e} \sim m_\mathrm{e}/m_\mathrm{i}$, electron rest mass inertia becomes non-negligible, and thus the HMHD approximation breaks down (Zocco & Schekochihin Reference Zocco and Schekochihin2011). To summarize, we assume the range of $\beta _\mathrm{e}$ and $\epsilon$ as

(2.6)\begin{equation} \epsilon \sim \frac{m_\mathrm{e}}{m_\mathrm{i}} \ll \beta_\mathrm{e} \ll \sqrt{\frac{m_\mathrm{e}}{m_\mathrm{i}}}. \end{equation}

We, then, follow the derivation of non-relativistic RMHD by Schekochihin et al. (Reference Schekochihin, Cowley, Dorland, Hammett, Howes, Quataert and Tatsuno2009). We explicitly keep $v_\mathrm{A}/c$ and $k_{\bot } d_\mathrm{i}$ in the ordering so that one can take the non-relativistic and/or long-wavelength limit simply by neglecting the corresponding terms. First, we adopt the expansion $\boldsymbol {u} = \boldsymbol {u}^{(1)} + \boldsymbol {u}^{(2)} + O(\epsilon ^3 v_\mathrm{A})$ and substitute it into (2.1a). The $O(\epsilon ^0 n_0 \omega )$ terms yield

(2.7)\begin{equation} \boldsymbol{\nabla}_{\bot}\boldsymbol{\cdot}\boldsymbol{u}_{\bot}^{(1)} = 0. \end{equation}

This allows us to use a stream function leading to $\boldsymbol {u}_{\bot }^{(1)} = \hat {\boldsymbol {z}}\times \boldsymbol {\nabla }_{\bot }\varPhi$, where $\varPhi = (c/B_0)\phi$, and $\phi$ is the electrostatic potential. From the $O(\epsilon ^1 n_0 \omega )$ terms, one obtains

(2.8)\begin{equation} \left( \frac{\partial }{\partial t} + \boldsymbol{u}_{\bot}^{(1)}\boldsymbol{\cdot}\boldsymbol{\nabla}_{\bot} \right)\frac{\delta n}{n_0} ={-}\left( \boldsymbol{\nabla}_{\bot}\boldsymbol{\cdot}\boldsymbol{u}_{\bot}^{(2)} + \frac{\partial u_{\|}^{(1)}}{\partial z} \right). \end{equation}

From the lowest-order terms in (2.1f), one can use a magnetic flux function leading to $v_\mathrm{A}(\delta \boldsymbol {B}_{\bot }^{(1)}/B_0) = \hat {\boldsymbol {z}}\times \boldsymbol {\nabla }_{\bot }\varPsi$, where $\varPsi = -(v_\mathrm{A}/B_0)A_{\|}$, and $A_{\|}$ is the parallel component of the vector potential. The $O(\epsilon ^0\omega )$ terms in (2.1b) give the pressure balance

(2.9)\begin{equation} \frac{B_0}{4{\rm \pi}}\delta B_{\|} ={-}\delta p_\mathrm{e} ={-}T_{\mathrm{e}0}\delta n. \end{equation}

Up to this point, the derivation is the same as the non-relativistic RMHD.

Next, we expand $\boldsymbol {E} = \boldsymbol {E}^{(1)} + \boldsymbol {E}^{(2)} + O(\epsilon ^3 v_{{\rm A}} B_0/c)$. Using electromagnetic potentials, the first- and second-order terms become

(2.10a)\begin{equation} \boldsymbol{E}^{(1)} ={-}\boldsymbol{\nabla}_{\bot}\phi, \quad \boldsymbol{E}^{(2)} ={-}\frac{1}{c}\frac{\partial \boldsymbol{A}}{\partial t} - \hat{\boldsymbol{z}}\frac{\partial \phi}{\partial z}. \end{equation}

We also expand $\boldsymbol {J} = \boldsymbol {J}^{(1)} + \boldsymbol {J}^{(2)} + O(\epsilon ^3 ck_{\bot }B_0)$, plug it into (2.1c) and remove $\rho _{\rm q}$ using (2.1d) to yield

(2.11) \begin{gather} \frac{c}{v_\mathrm{A} B_0}\left( \underbrace{\boldsymbol{E}^{(1)}}_{{\sim}\epsilon} + \underbrace{\boldsymbol{E}^{(2)}}_{{\sim}\epsilon^2} \right) + \frac{c}{v_\mathrm{A}}\left( \underbrace{\frac{\delta n}{n_0}}_{{\sim}\epsilon^2} - \underbrace{\frac{\boldsymbol{\nabla}_{\bot}\boldsymbol{\cdot}\boldsymbol{E}_{\bot}^{(1)}}{4{\rm \pi} e n_0}}_{{\sim}\epsilon^2(k_{\bot}d_{{\rm i}})(v_{{\rm A}}/c)^2} \right)\frac{\boldsymbol{E}^{(1)}}{B_0} + \frac{1}{v_{{\rm A}}}\left( \underbrace{\boldsymbol{u}^{(1)}}_{{\sim}\epsilon} - \underbrace{\frac{\boldsymbol{J}^{(1)}}{en_0}}_{{\sim}\epsilon (k_{\bot} d_{{\rm i}})} \right)\times\hat{\boldsymbol{z}} \nonumber\\ +\, \frac{1}{v_{{\rm A}}}\left( \underbrace{\boldsymbol{u}^{(1)}}_{{\sim}\epsilon^2} - \underbrace{\frac{\boldsymbol{J}^{(1)}}{en_0}}_{{\sim}\epsilon^2 (k_{\bot} d_{{\rm i}})} \right)\times\frac{\delta \boldsymbol{B}}{B_0} + \frac{1}{v_{{\rm A}}}\left[ \underbrace{\left( \frac{\delta n}{n_0} \right)\boldsymbol{u}^{(1)} + \boldsymbol{u}^{(2)}}_{{\sim}\epsilon^2} + \underbrace{\left( \frac{\delta n}{n_0} \right)\frac{\boldsymbol{J}^{(1)}}{en_0} - \frac{\boldsymbol{J}^{(2)}}{en_0}}_{{\sim}\epsilon^2 (k_{\bot} d_{{\rm i}})} \right]\times\hat{\boldsymbol{z}} \nonumber\\ +\, \frac{c}{v_{{\rm A}} en_0 B_0}\left[\underbrace{\boldsymbol{\nabla}_{\bot}\delta p_{{\rm e}}}_{{\sim}\epsilon\beta_{{\rm e}} (k_{\bot}d_{{\rm i}})} + \underbrace{\left( \frac{\delta n}{n_0} \right)\boldsymbol{\nabla}_{\bot}\delta p_{{\rm e}} + \hat{\boldsymbol{z}}\frac{\partial }{\partial z}\delta p_{{\rm e}}}_{{\sim}\epsilon^2\beta_{{\rm e}} (k_{\bot}d_{{\rm i}})}\right] = 0, \end{gather}

where $d_{{\rm i}} = \sqrt {m_{{\rm i}} c^2/4{\rm \pi} n_0 e^2}$ is the ion inertial length. As we mentioned above, $\beta _{{\rm e}}$ is assumed to be $O(\epsilon ^0)$ because $\epsilon \ll \beta _{{\rm e}}$. We then manipulate (2.1e) to get

(2.12)\begin{align} \frac{4{\rm \pi}\boldsymbol{J}}{cB_0} & =-\frac{1}{c^2}\hat{\boldsymbol{z}}\times\frac{\partial }{\partial t}\left( \underbrace{\boldsymbol{u}_{\bot}^{(1)}}_{{\sim} \epsilon^2 k_{\bot}(v_{{\rm A}}/c)^2} - \underbrace{\frac{\boldsymbol{J}_{\bot}^{(1)}}{en_0}}_{{\sim} \epsilon^2 k_{\bot}(k_{\bot} d_{{\rm i}})(v_{{\rm A}}/c)^2} \right) + \underbrace{\frac{1}{ce n_0 B_0}\frac{\partial }{\partial t}\boldsymbol{\nabla}_{\bot}\delta p_{{\rm e}}}_{{\sim} \epsilon^2 \beta_{{\rm e}} k_{\bot}(k_{\bot} d_{{\rm i}})(v_{{\rm A}}/c)^2} \nonumber\\ & \quad +\, \underbrace{\hat{\boldsymbol{z}}\times\frac{\partial }{\partial z}\frac{\delta \boldsymbol{B}_{\bot}}{B_0}}_{{\sim} \epsilon^2 k_{\bot}} + \underbrace{\boldsymbol{\nabla}_{\bot}\left( \frac{\delta B_{\|}}{B_0} \right)\times\hat{\boldsymbol{z}}}_{{\sim} \epsilon k_{\bot}} + \underbrace{\boldsymbol{\nabla}_{\bot} \times\frac{\delta\boldsymbol{B}_{\bot}}{B_0}}_{{\sim} \epsilon k_{\bot}} + O(\epsilon^3 k_{\bot}). \end{align}

Collecting the terms order-by-order, one obtains

(2.13) \begin{equation} \begin{gathered} \boldsymbol{J}_{\bot}^{(1)} = \frac{c}{4{\rm \pi}}\boldsymbol{\nabla}_{\bot} \delta B_{\|} \times \hat{\boldsymbol{z}},\quad J_{\|}^{(1)} = \frac{c}{4{\rm \pi}}\hat{\boldsymbol{z}}\boldsymbol{\cdot}(\boldsymbol{\nabla}_{\bot}\times\delta\boldsymbol{B}_{\bot}),\\ \boldsymbol{J}_{\bot}^{(2)} = \frac{c}{4{\rm \pi}}\left[ -\frac{B_0}{c^2}\hat{\boldsymbol{z}}\times\frac{\partial }{\partial t}\left( \boldsymbol{u}_{\bot}^{(1)} - \frac{\boldsymbol{J}_{\bot}^{(1)}}{en_0} \right) + \frac{1}{cen_0}\frac{\partial }{\partial t}\boldsymbol{\nabla}_{\bot}\delta p_{{\rm e}} + \hat{\boldsymbol{z}}\times\frac{\partial }{\partial z}\delta\boldsymbol{B}_{\bot} \right], \quad J_{\|}^{(2)} = 0. \end{gathered} \end{equation}

The terms including $\partial /\partial t$ in $\boldsymbol {J}_{\bot }^{(2)}$ originated from the displacement current, which disappears in the non-relativistic limit, and all the other terms are the same as the non-relativistic case. Plugging $\boldsymbol {J}_{\bot }^{(1)}$ into the $O(\epsilon )$ terms in (2.11), one obtains the pressure balance (2.9) again, where the isothermal electrons are assumed. Then, we can further manipulate $\boldsymbol {J}_{\bot }^{(2)}$ as

(2.14)\begin{equation} \boldsymbol{J}_{\bot}^{(2)} = \frac{c}{4{\rm \pi}}\left( \frac{B_0}{c^2}\frac{\partial }{\partial t}\boldsymbol{\nabla}_{\bot}\varPhi + \hat{\boldsymbol{z}}\times\frac{\partial }{\partial z}\delta\boldsymbol{B}_{\bot} \right). \end{equation}

This is notable because the Hall term and the electron pressure gradient term, both of which becomes finite at the $d_{{\rm i}}$ scale, are cancelled, and therefore, the electric current (up to the second order) is identical to that of relativistic RMHD. The reason for this cancellation is rather straightforward. We evaluate the electric field $\boldsymbol {E}$ in the displacement current by using Ohm's law, which is the momentum equation of electrons. Therefore, the pressure balance cancels some of the terms of the electric current exactly in the same way as the first order of the momentum equation in the relativistic RMHD (which is identical to (2.16) shown below).

Next, we substitute (2.10a,b) and (2.13) into (2.1b) to get the equations for $\varPhi$ and $u_{\|}$. As we found in the previous paragraph, both $\boldsymbol {E}$ and $\boldsymbol {J}$ do not contain corrections due to the Hall effect, meaning that (2.1b) ends up with the same equations as the relativistic RMHD. The $z$ component of $O(\epsilon ^1\omega )$ terms give

(2.15)\begin{equation} \frac{\partial u_{\|}}{\partial t} + \left\{ \varPhi, u_{\|} \right\} = (1 + \sigma)v_{{\rm A}}^2\left[ \frac{\partial }{\partial z}\left( \frac{\delta B_{\|}}{B_0} \right) + \frac{1}{v_{{\rm A}}}\left\{ \varPsi, \frac{\delta B_{\|}}{B_0}\right\} \right], \end{equation}

while multiplying $\hat {\boldsymbol {z}}\boldsymbol {\cdot }\boldsymbol {\nabla }_{\bot }\times$ to $O(\epsilon ^1\omega )$ terms gives

(2.16)\begin{equation} \frac{\partial }{\partial t}\boldsymbol{\nabla}_{\bot}^2\varPhi + \left\{ \varPhi, \boldsymbol{\nabla}_{\bot}^2\varPhi \right\} = v_{{\rm A}}\left( \frac{\partial }{\partial z}\boldsymbol{\nabla}_{\bot}^2\varPsi + \frac{1}{v_{{\rm A}}}\left\{ \varPsi, \boldsymbol{\nabla}_{\bot}^2\varPsi \right\} \right), \end{equation}

where $\left \{ f, g \right \} = (\partial _x f)(\partial _y g) - (\partial _x g)(\partial _y f)$.

Next, we derive the equations for $\varPsi$ and $\delta B_{\|}$. This time, unlike the momentum equation, there will be the corrections due to the Hall effect and electron pressure gradient. The $O(\epsilon ^2)$ terms in (2.11) are

(2.17)\begin{gather} \frac{\boldsymbol{E}^{(2)}}{B_0} + \left(\frac{\delta n}{n_0} - \frac{\boldsymbol{\nabla}_{\bot}\boldsymbol{\cdot}\boldsymbol{E}_{\bot}^{(1)}}{4{\rm \pi} e n_0} \right)\frac{\boldsymbol{E}^{(1)}}{B_0} + \frac{1}{c}\left( \boldsymbol{u}^{(1)} - \frac{\boldsymbol{J}^{(1)}}{en_0} \right)\times\frac{\delta \boldsymbol{B}}{B_0} \nonumber\\ +\, \frac{1}{c}\left[ \left( \frac{\delta n}{n_0} \right)\boldsymbol{u}^{(1)} + \boldsymbol{u}^{(2)} - \frac{\boldsymbol{J}^{(2)}}{en_0} \right]\times\hat{\boldsymbol{z}} + \frac{1}{en_0 B_0}\left(\hat{\boldsymbol{z}}\frac{\partial }{\partial z}\delta p_{{\rm e}}\right) = 0. \end{gather}

Note that a few terms have been cancelled out using the pressure balance (2.9). The $z$-component of (2.17) yields

(2.18)\begin{equation} \frac{\partial \varPsi}{\partial t} = v_{{\rm A}}\left[ \frac{\partial }{\partial z}\left( \varPhi + \sqrt{(1 + \sigma)}\, d_{{\rm i}} v_{{\rm A}}\frac{\delta B_{\|}}{B_0} \right) + \frac{1}{v_{{\rm A}}}\left\{ \varPsi,\, \varPhi + \sqrt{(1 + \sigma)}\, d_{{\rm i}} v_{{\rm A}}\frac{\delta B_{\|}}{B_0} \right\} \right]. \end{equation}

Meanwhile, (2.8) and (2.9) are rearranged as

(2.19)\begin{equation} \boldsymbol{\nabla}_{\bot}\boldsymbol{\cdot}\boldsymbol{u}_{\bot}^{(2)} + \frac{\partial u_{\|}^{(1)}}{\partial z} = \frac{\sigma c^2}{c_{{\rm S}}^2}\frac{\mathrm{d}}{\mathrm{d}t}\left( \frac{\delta B_{\|}}{B_0} \right), \end{equation}

where $c_{{\rm S}} = \sqrt {(n_{{{\rm i}}0}T_{{{\rm e}}0})/(n_{{{\rm e}}0}m_{\rm i})} \ll (m_{{\rm e}}/m_{{\rm i}})^{1/4}c$ is the sound speed, and (2.13) and (2.14) are rearranged as

(2.20)\begin{equation} \frac{1}{en_0}\left( \boldsymbol{\nabla}_{\bot}\boldsymbol{\cdot}\boldsymbol{J}_{\bot}^{(2)} + \frac{\partial J_{\|}^{(1)}}{\partial z} \right) = \frac{\sqrt{(1 + \sigma)}\, d_{{\rm i}} v_{{\rm A}}}{c^2}\frac{\partial }{\partial t}\boldsymbol{\nabla}_{\bot}^2\varPhi. \end{equation}

Note that the right-hand side of (2.20) is equal to zero in the non-relativistic limit (i.e. $\boldsymbol {\nabla }\boldsymbol {\cdot }\boldsymbol {J} = 0$). Combining (2.19) and (2.20) with $\hat {\boldsymbol {z}}\boldsymbol {\cdot }\boldsymbol {\nabla }_{\bot }\times$ to (2.17) yields

(2.21)\begin{gather} -\left( 1 + \frac{\sigma c^2}{c_{{\rm S}}^2} \right)\left[ \frac{\partial }{\partial t}\left( \frac{\delta B_{\|}}{B_0} \right) + \left\{ \varPhi,\, \frac{\delta B_{\|}}{B_0} \right\} \right] + \frac{\sqrt{(1 + \sigma)}\, d_{{\rm i}} v_{{\rm A}}}{c^2}\left( \frac{\partial }{\partial t}\boldsymbol{\nabla}_{\bot}^2\varPhi + \left\{ \varPhi,\, \boldsymbol{\nabla}_{\bot}^2\varPhi \right\} \right) \nonumber\\ +\, \frac{\partial u_{\|}}{\partial z} + \frac{1}{v_{{\rm A}}}\left\{ \varPsi,\, u_{\|} \right\} - \sqrt{(1 + \sigma)}\, d_{{\rm i}} \left( \frac{\partial }{\partial z}\boldsymbol{\nabla}_{\bot}^2\varPsi + \frac{1}{v_{{\rm A}}}\left\{ \varPsi,\, \boldsymbol{\nabla}_{\bot}^2\varPsi \right\} \right) = 0. \end{gather}

The time derivative of $\boldsymbol {\nabla }_{\bot }^2\varPhi$ can be eliminated by substituting (2.16), which results in

(2.22)\begin{gather} \left( 1 + \frac{\sigma c^2}{c_{{\rm S}}^2} \right)\left[ \frac{\partial }{\partial t}\left( \frac{\delta B_{\|}}{B_0} \right) + \left\{ \varPhi,\, \frac{\delta B_{\|}}{B_0} \right\} \right] \nonumber\\ = \frac{\partial u_{\|}}{\partial z} + \frac{1}{v_{{\rm A}}}\left\{ \varPsi,\, u_{\|} \right\} - \sqrt{(1 + \sigma)}\, d_{{\rm i}}\left( 1 - \frac{v_{{\rm A}}^2}{c^2} \right)\left( \frac{\partial }{\partial z} + \frac{1}{v_{{\rm A}}^2}\left\{ \varPsi,\, \boldsymbol{\nabla}_{\bot}^2\varPsi \right\} \right). \end{gather}

Finally, (2.15), (2.16), (2.18) and (2.22) are reorganized into

(2.23a)\begin{gather} \frac{\mathrm{d}}{\mathrm{d}t}\boldsymbol{\nabla}_{\bot}^2\varPhi = v_{{\rm A}}\boldsymbol{\nabla}_{\|}\boldsymbol{\nabla}_{\bot}^2\varPsi, \end{gather}
(2.23b)\begin{gather}\frac{\partial \varPsi}{\partial t} = v_{{\rm A}}\boldsymbol{\nabla}_{\|}\left( \varPhi + v_{{\rm A}}\rho_{{\rm H}}\mathcal{B} \right), \end{gather}
(2.23c)\begin{gather}\frac{\mathrm{d}\mathcal{U}}{\mathrm{d}t} = v_{{\rm S}}\boldsymbol{\nabla}_{\|}\mathcal{B}, \end{gather}
(2.23d)\begin{gather}\frac{\mathrm{d}\mathcal{B}}{\mathrm{d}t} = \boldsymbol{\nabla}_{\|}\left( v_{{\rm S}}\mathcal{U} - \rho_{{\rm H}} \boldsymbol{\nabla}_{\bot}^2\varPsi \right), \end{gather}

where

(2.24a)\begin{gather} \frac{\mathrm{d}}{\mathrm{d}t} = \frac{\partial }{\partial t} + \boldsymbol{u}_{\bot}\boldsymbol{\cdot}\boldsymbol{\nabla}_{\bot} = \frac{\partial }{\partial t} + \left\{ \varPhi, \ldots \right\},\quad \boldsymbol{\nabla}_{\|} = \frac{\partial }{\partial z} + \frac{\delta\boldsymbol{B}_{\bot}}{B_0}\boldsymbol{\cdot}\boldsymbol{\nabla}_{\bot} = \frac{\partial }{\partial z} + \frac{1}{v_{{\rm A}}} \left\{ \varPsi, \ldots \right\}, \end{gather}
(2.24b)\begin{gather}\mathcal{U} = \frac{u_{\|}}{v_{{\rm A}}}, \quad \mathcal{B} = \sqrt{(1 + \sigma)\left[ 1 + (1 + \sigma)v_{{\rm A}}^2/c_{{\rm S}}^2 \right]}\,\frac{\delta B_{\|}}{B_0}, \end{gather}
(2.24c) \begin{gather}v_{{\rm A}} = \sqrt{\frac{\sigma}{1 + \sigma}}\,c, \quad v_{{\rm S}} = \sqrt{\frac{\sigma}{1 + (1 + \sigma)v_{{\rm A}}^2/c_{{\rm S}}^2}}\,c, \quad \rho_{H} = \sqrt{\frac{1}{1 + (1 + \sigma)v_{{\rm A}}^2/c_{{\rm S}}^2}}\,d_{{\rm i}}. \end{gather}

Note that $v_{{\rm A}}^2/c_{{\rm S}}^2 \approx 2/\beta _{{\rm e}}$ when $T_{{{\rm i}}0}/T_{{{\rm e}}0} \ll 1$. We find that (2.23a)–(2.23d) are identical to non-relativistic HRMHD ((5.14)–(5.17) of S19), except for the definition of constants (2.24c) which become those of S19 in the limit of $\sigma \to 0$. Thereby, all the physical properties of the non-relativistic HRMHD (e.g. the conservation of energy and helicity ((5.24) and (5.68) of S19) and the linear dispersion relation ((5.26) of S19) are valid even in a magnetically dominated regime $\sigma \gtrsim 1$. Meanwhile, one retrieves the relativistic RMHD by taking the limit of $\rho _{{\rm H}} \to 0$. Note that RMHD of Chandran et al. (Reference Chandran, Foucart and Tchekhovskoy2018) and TenBarge et al. (Reference TenBarge, Ripperda, Chernoglazov, Bhattacharjee, Mahlmann, Most, Juno, Yuan and Philippov2021) are written in the Elsässer variables while (2.23a)–(2.23d) are not; as is shown by Galtier (Reference Galtier2006) that introducing the Elsässer for HMHD is complicated.

There are three important facts that make relativistic HRMHD and non-relativistic HRMHD formally identical. The first is that the electric current of relativistic HRMHD is identical to that of relativistic RMHD due to the electron pressure balance (see (2.13)). The second is that the displacement current only appears in the second order of the electric current. The third is that the Lorentz force due to the displacement current (i.e. $\boldsymbol {J}^{(2)}\times \boldsymbol {B}_0$) is the time derivative of $\boldsymbol {E}^{(1)}\times \boldsymbol {B}_0$ drift, which happens to be the left-hand side of the momentum equation (2.1b) and balances the non-relativistic Lorentz force $\boldsymbol {J}^{(1)}\times \delta \boldsymbol {B}$ in the right-hand side (given the pressure balance). Thus, the relativistic Lorentz force ends up with merely proportional to the non-relativistic Lorentz force, and its prefactor does not include $d_{{\rm i}}$.

Lastly, one finds that the Hall transition scale $\rho _{{\rm H}}$ becomes smaller and vanishes eventually as the magnetization $\sigma$ increases, meaning that the fluctuations behave like those of RMHD at $d_{{\rm i}}$ scale when the magnetization is strong. This behaviour was first discovered by Kawazura (Reference Kawazura2017), and the reason for it is that the Hall transition happens at the scale of $\sim v_{{\rm A}}/\varOmega _{{\rm i}}$ (which is equal to $d_{{\rm i}}$ in the non-relativistic limit) where $\varOmega _{{\rm i}}$ is ion cyclotron frequency, and this scale becomes smaller as the magnetization becomes larger because $v_{{\rm A}} \to c$ and $\varOmega _{{\rm i}} \to \infty$ as $\sigma \to \infty$.

3. Discussion

In this paper, we have shown that the properties of fluctuations in the sub-Hall transition scale ($k_{\bot } \rho _{{\rm H}} \gtrsim 1$) found by non-relativistic HRMHD remain the same even when the mean magnetic field is relativistically strong and/or the electron temperature is relativistically hot (but much less than $(m_{{\rm i}}/m_{{\rm e}})^{1/2}m_{{\rm e}} c^2$). More specifically, for example, S19 theoretically showed that the Alfvénic and compressive cascades in the RMHD range ($k_{\bot } \rho _{{\rm H}} \ll 1$) are rearranged into KAW and ICW cascades below the Hall transition scale ($k_{\bot } \rho _{{\rm H}} \gg 1$), and the scalings of KAW cascade are $k_{\bot }^{-7/3}$ for the magnetic energy and $k_{\bot }^{-13/3}$ for the kinetic energy while those of ICW cascade are $k_{\bot }^{-5/3}$ for the magnetic energy and $k_{\bot }^{-11/3}$ for the kinetic energy. These scalings are consistent with the numerical simulation of non-relativistic HMHD (Meyrand & Galtier Reference Meyrand and Galtier2012). We find that the same scalings apply to the relativistic regime, although there have been no simulation of turbulence in relativistic ion and electron plasmas that elucidated the scalings at the transition scale. Note that the relativistic PIC simulation of ion and electron plasma turbulence found the presence of the spectral break at the ion Larmor scale (Zhdankin et al. Reference Zhdankin, Uzdensky, Werner and Begelman2019). Furthermore, S19 also showed that the KAW and ICW cascades eventually turn into electron and ion heating, respectively, and we have shown that this scenario of energy partition between ions and electrons is also true in the relativistic regime.

In closing, we admit that the ordering assumptions we have made in this study, namely (2.4) and (2.6), are rather restrictive and may not be directly relevant to realistic astrophysical objects. However, even though HMHD is consistent with kinetic theory only for the cold ion limit (Ito et al. Reference Ito, Hirose, Mahajan and Ohsaki2004; Howes Reference Howes2009), the simulations of turbulence at the transition scale via HMHD and through the hybrid PIC with $T_{{\rm i}} = T_{{\rm e}}$ demonstrated nearly the same results (Papini et al. Reference Papini, Franci, Landi, Verdini, Matteini and Hellinger2019), suggesting that the predictions made by HMHD are practically useful beyond its theoretical limitations. We think the same holds true for our relativistic HRMHD.

Acknowledgements

The author thanks S. Kimura for fruitful conversations.

Editor D. Uzdensky thanks the referees for their advice in evaluating this article.

Funding

This work was supported by JSPS KAKENHI grant JP20K14509.

Declaration of interests

The author reports no conflict of interest.

Footnotes

1 Note that the non-relativistic version of XMHD (Kimura & Morrison Reference Kimura and Morrison2014; Abdelhamid, Kawazura & Yoshida Reference Abdelhamid, Kawazura and Yoshida2015) was used for turbulence of solar wind (Abdelhamid, Lingam & Mahajan Reference Abdelhamid, Lingam and Mahajan2016).

References

Abdelhamid, H.M., Kawazura, Y. & Yoshida, Z. 2015 Hamiltonian formalism of extended magnetohydrodynamics. J. Phys. A: Math. Theor. 48 (23), 235502.CrossRefGoogle Scholar
Abdelhamid, H.M., Lingam, M. & Mahajan, S.M. 2016 Extended MHD turbulence and its applications to the solar wind. Astrophys. J. 829 (2), 87.CrossRefGoogle Scholar
Asenjo, F.A. & Comisso, L. 2015 Generalized magnetofluid connections in relativistic magnetohydro-dynamics. Phys. Rev. Lett. 114 (11), 115003.CrossRefGoogle Scholar
Asenjo, F.A. & Comisso, L. 2019 Gravitational electromotive force in magnetic reconnection around schwarzschild black holes. Phys. Rev. D 99 (6), 063017.CrossRefGoogle Scholar
Asenjo, F.A., Comisso, L. & Mahajan, S.M. 2015 Generalized magnetofluid connections in pair plasmas. Phys. Plasmas 22 (12), 122109.CrossRefGoogle Scholar
Chandran, B.D.G., Foucart, F. & Tchekhovskoy, A. 2018 Heating of accretion-disk coronae and jets by general relativistic magnetohydrodynamic turbulence. J. Plasma Phys. 84 (3), 905840310.CrossRefGoogle Scholar
Chen, C.H.K., Mallet, A., Yousef, T.A., Schekochihin, A.A. & Horbury, T.S. 2011 Anisotropy of Alfvénic turbulence in the solar wind and numerical simulations. Mon. Not. R. Astron. Soc. 415 (4), 32193226.CrossRefGoogle Scholar
Comisso, L. & Asenjo, F.A. 2014 Thermal-inertial effects on magnetic reconnection in relativistic pair plasmas. Phys. Rev. Lett. 113 (4), 045001.CrossRefGoogle ScholarPubMed
Comisso, L. & Asenjo, F.A. 2018 Collisionless magnetic reconnection in curved spacetime and the effect of black hole rotation. Phys. Rev. D 97 (4), 043007.CrossRefGoogle Scholar
Comisso, L. & Asenjo, F.A. 2020 Generalized magnetofluid connections in a curved spacetime. Phys. Rev. D 102 (2), 023032.CrossRefGoogle Scholar
Comisso, L. & Sironi, L. 2018 Particle acceleration in relativistic plasma turbulence. Phys. Rev. Lett. 121 (25), 255101.CrossRefGoogle ScholarPubMed
Foucart, F., Chandra, M., Gammie, C.F. & Quataert, E. 2016 Evolution of accretion discs around a kerr black hole using extended magnetohydrodynamics. Mon. Not. R. Astron. Soc. 456 (2), 13321345.CrossRefGoogle Scholar
Foucart, F., Chandra, M., Gammie, C.F., Quataert, E. & Tchekhovskoy, A. 2017 How important is non-ideal physics in simulations of sub-Eddington accretion on to spinning black holes? Mon. Not. R. Astron. Soc. 470 (2), 22402252.CrossRefGoogle Scholar
Galtier, S. 2006 Wave turbulence in incompressible Hall magnetohydrodynamics. J. Plasma Phys. 72 (5), 721769.CrossRefGoogle Scholar
Gómez, D.O., Mahajan, S.M. & Dmitruk, P. 2008 Hall magnetohydrodynamics in a strong magnetic field. Phys. Plasmas 15 (10), 102303.CrossRefGoogle Scholar
Howes, G.G. 2009 Limitations of Hall MHD as a model for turbulence in weakly collisional plasmas. Nonlinear Process. Geophys. 16 (2), 219232.CrossRefGoogle Scholar
Howes, G.G., Cowley, S.C., Dorland, W., Hammett, G.W., Quataert, E. & Schekochihin, A.A. 2008 A model of turbulence in magnetized plasmas: implications for the dissipation range in the solar wind. J. Geophys. Res. 113 (A5), A05103.Google Scholar
Ito, A., Hirose, A., Mahajan, S.M. & Ohsaki, S. 2004 Remarks on the discrete Alfvén wave spectrum induced by the Hall current. Phys. Plasmas 11 (12), 56435647.CrossRefGoogle Scholar
Kadomtsev, B.B. & Pogutse, O.P. 1974 Nonlinear helical perturbations of a plasma in the tokamak. Sov. Phys. JETP 38, 283.Google Scholar
Kawazura, Y. 2017 Modification of magnetohydrodynamic waves by the relativistic Hall effect. Phys. Rev. E 96 (1), 013207.CrossRefGoogle ScholarPubMed
Kawazura, Y. 2022 Yet another modification of relativistic magnetohydrodynamic waves: electron thermal inertia. Phys. Lett. A 443, 128199.CrossRefGoogle Scholar
Kawazura, Y., Miloshevich, G. & Morrison, P.J. 2017 Action principles for relativistic extended magnetohydrodynamics: a unified theory of magnetofluid models. Phys. Plasmas 24 (2), 022103.CrossRefGoogle Scholar
Kawazura, Y., Schekochihin, A.A., Barnes, M., Dorland, W. & Balbus, S.A. 2022 Energy partition between Alfvénic and compressive fluctuations in magnetorotational turbulence with near-azimuthal mean magnetic field. J. Plasma Phys. 88 (3), 905880311.CrossRefGoogle Scholar
Kimura, K. & Morrison, P.J. 2014 On energy conservation in extended magnetohydrodynamics. Phys. Plasmas 21 (8), 082101.CrossRefGoogle Scholar
Koide, S. 2009 Generalized relativistic magnetohydrodynamic equations for pair and electron–ion plasmas. Astrophys. J. 696 (2), 22202233.CrossRefGoogle Scholar
Kraichnan, R.H. 1965 Inertial-range spectrum of hydromagnetic turbulence. Phys. Fluids 8, 1385.CrossRefGoogle Scholar
Meyrand, R. & Galtier, S. 2012 Spontaneous chiral symmetry breaking of Hall magnetohydrodynamic turbulence. Phys. Rev. Lett. 109 (19), 194501.CrossRefGoogle ScholarPubMed
Nättilä, J. & Beloborodov, A.M. 2022 Heating of magnetically dominated plasma by Alfvén-wave turbulence. Phys. Rev. Lett. 128 (7), 075101.CrossRefGoogle ScholarPubMed
Papini, E., Franci, L., Landi, S., Verdini, A., Matteini, L. & Hellinger, P. 2019 Can hall magnetohydrodynamics explain plasma turbulence at sub-ion scales? Astrophys. J. 870 (1), 52.CrossRefGoogle Scholar
Ripperda, B., Bacchini, F. & Philippov, A.A. 2020 Magnetic reconnection and hot spot formation in black hole accretion disks. Astrophys. J. 900 (2), 100.CrossRefGoogle Scholar
Ripperda, B., Liska, M., Chatterjee, K., Musoke, G., Philippov, A.A., Markoff, S.B., Tchekhovskoy, A. & Younsi, Z. 2022 Black hole flares: ejection of accreted magnetic flux through 3D plasmoid-mediated reconnection. Astrophys. J. Lett. 924 (2), L32.CrossRefGoogle Scholar
Schekochihin, A.A., Cowley, S.C., Dorland, W., Hammett, G.W., Howes, G.G., Quataert, E. & Tatsuno, T. 2009 Astrophysical gyrokinetics: kinetic and fluid turbulent cascades in magnetized weakly collisional plasmas. Astrophys. J. Supp. 182, 310.CrossRefGoogle Scholar
Schekochihin, A.A., Kawazura, Y. & Barnes, M.A. 2019 Constraints on ion versus electron heating by plasma turbulence at low beta. J. Plasma Phys. 85, 905850303.CrossRefGoogle Scholar
Strauss, H.R. 1976 Nonlinear, three-dimensional magnetohydrodynamics of noncircular tokamaks. Phys. Fluids 19, 134.CrossRefGoogle Scholar
TenBarge, J.M., Ripperda, B., Chernoglazov, A., Bhattacharjee, A., Mahlmann, J.F., Most, E.R., Juno, J., Yuan, Y. & Philippov, A.A. 2021 Weak Alfvénic turbulence in relativistic plasmas. Part 1. Dynamical equations and basic dynamics of interacting resonant triads. J. Plasma Phys. 87 (6), 905870614.CrossRefGoogle Scholar
Watanabe, T.H. 2010 Feedback instability in the magnetosphere-ionosphere coupling system: revisited. Phys. Plasmas 17 (2), 022904.CrossRefGoogle Scholar
Yang, S.D. 2019 a Relativistic sweet-parker current sheet and its macroscopic evolution. Astrophys. J. 882 (2), 101.CrossRefGoogle Scholar
Yang, S.D. 2019 b Relativistic tearing mode in pair plasmas and application to magnetic giant flares. Astrophys. J. 880 (1), 44.CrossRefGoogle Scholar
Zhdankin, V., Uzdensky, D.A., Werner, G.R. & Begelman, M.C. 2019 Electron and ion energization in relativistic plasma turbulence. Phys. Rev. Lett. 122, 055101.CrossRefGoogle ScholarPubMed
Zhdankin, V., Walker, J., Boldyrev, S. & Lesur, G. 2017 Kinetic turbulence in relativistic plasma: from thermal bath to nonthermal continuum. Phys. Rev. Lett. 118 (5), 055103.CrossRefGoogle ScholarPubMed
Zocco, A. & Schekochihin, A.A. 2011 Reduced fluid-kinetic equations for low-frequency dynamics, magnetic reconnection, and electron heating in low-beta plasmas. Phys. Plasmas 18 (10), 102309.CrossRefGoogle Scholar