Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-20T01:47:51.981Z Has data issue: false hasContentIssue false

‘Effective’ collisions in weakly magnetized collisionless plasma: importance of Pitaevski’s effect for magnetic reconnection

Published online by Cambridge University Press:  09 February 2016

Lev M. Zelenyi
Affiliation:
Space Research Institute (IKI), 84/32 Profsoyuznaya Str, Moscow 117997, Russia
Anton V. Artemyev*
Affiliation:
Space Research Institute (IKI), 84/32 Profsoyuznaya Str, Moscow 117997, Russia Department of Earth, Planetary, and Space Sciences and Institute of Geophysics and Planetary Physics, University of California, 603 Charles E Young Dr E, Los Angeles, CA 90095, USA
*
Email address for correspondence: ante0226@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

In this paper we revisit the paradigm of space science turbulent dissipation traditionally considered as myth (Coroniti, Space Sci. Rev., vol. 42, 1985, pp. 399–410). We demonstrate that due to approach introduced by Pitaevskii (Sov. J. Expl Theor. Phys., vol. 44, 1963, pp. 969–979; (in Russian)) (the effect of a finite Larmor radius on a classical collision integral) dissipation induced by effective interaction with microturbulence produces a significant effect on plasma dynamics, especially in the vicinity of the reconnection region. We estimate the multiplication factor of collision frequency in the collision integral for short wavelength perturbations. For waves propagating transverse to the background magnetic field, this factor is approximately $({\it\rho}_{e}k_{x})^{2}$ with ${\it\rho}_{e}$ an electron gyroradius and where $k_{x}$ is a transverse wavenumber. We consider recent spacecraft observations in the Earth’s magnetotail reconnection region to the estimate possible impact of this multiplication factor. For small-scale reconnection regions this factor can significantly increase the effective collision frequency produced both by lower-hybrid drift turbulence and by kinetic Alfvén waves. We discuss the possibility that the Pitaevskii’s effect may be responsible for the excitation of a resistive electron tearing mode in thin current sheets formed in the outflow region of the primary X-line.

Type
Research Article
Copyright
© Cambridge University Press 2016 

1. Introduction

There are three main triggering processes for magnetic reconnection in equilibrium collisionless plasma systems: Landau resonance of tearing mode perturbations and demagnetized ions (e.g. Schindler Reference Schindler1974; Galeev & Zelenyi Reference Galeev and Zelenyi1976), inertia of magnetized electrons (e.g. Laval, Pellat & Vuillemin Reference Laval, Pellat and Vuillemin1966; Porcelli et al. Reference Porcelli, Borgogno, Califano, Grasso, Ottaviani and Pegoraro2002; Zelenyi & Artemyev Reference Zelenyi and Artemyev2013, and references therein) and effective collisions induced by particle scattering on electromagnetic turbulence (e.g. Huba, Gladd & Papadopoulos Reference Huba, Gladd and Papadopoulos1977; Coroniti Reference Coroniti1980; Büchner & Zelenyi Reference Büchner and Zelenyi1987). All these three processes can drive the instability of an equilibrium (or quasi-equilibrium) current sheet and result in magnetic field reconfiguration. The situation can be much more complicated for dynamical plasma systems where the turbulent magnetic reconnection destroys (or reforms) current sheets and magnetic discontinuities (Servidio et al. Reference Servidio, Dmitruk, Greco, Wan, Donato, Cassak, Shay, Carbone and Matthaeus2011; Greco et al. Reference Greco, Valentini, Servidio and Matthaeus2012; Rappazzo et al. Reference Rappazzo, Matthaeus, Ruffolo, Servidio and Velli2012; Karimabadi et al. Reference Karimabadi, Roytershteyn, Daughton and Liu2013). However, in this study we concentrate on consideration of small perturbations of an equilibrium current sheet. This is typical situation for magnetic reconnection in the Earth’s magnetotail where the ion Landau resonance is considered to be a main candidate for tearing mode excitation (Schindler Reference Schindler2006; Zelenyi et al. Reference Zelenyi, Artemyev, Malova and Popov2008; Sitnov & Swisdak Reference Sitnov and Swisdak2011) where the numerous spacecraft observations have revealed the manifestations of reconnection events (e.g. Petrukovich et al. Reference Petrukovich, Sergeev, Zelenyi, Mukai, Yamamoto, Kokubun, Shiokawa, Deehr, Budnick and Büchner1998; Baumjohann Reference Baumjohann2002; Angelopoulos et al. Reference Angelopoulos, McFadden, Larson, Carlson, Mende, Frey, Phan, Sibeck, Glassmeier and Auster2008; Nagai et al. Reference Nagai, Shinohara, Fujimoto, Matsuoka, Saito and Mukai2011). However, many physical problems of ion (resonant) tearing mode still remain unsolved (see discussion in Pellat, Coroniti & Pritchett Reference Pellat, Coroniti and Pritchett1991; Quest, Karimabadi & Brittnacher Reference Quest, Karimabadi and Brittnacher1996; Sitnov et al. Reference Sitnov, Sharma, Guzdar and Yoon2002). Thus, the idea of effective dissipation induced by particle scattering due to wave–particle interaction appears to be quite promising. Previous estimates (Coroniti Reference Coroniti1985) and modern spacecraft observations (e.g. Eastwood et al. Reference Eastwood, Phan, Bale and Tjulin2009) were somewhat pessimistic concerning this idea – the observed level of electromagnetic turbulence appears to be insufficient to support the required effective collision conductivity in the Earth’s magnetotail. In this paper, we consider an additional effect responsible for the enhancement of effective collisions. Although, we concentrate on magnetotail reconnection, the proposed effect can also be important for laboratory devices where a weakly collisional reconnection geometry is reproduced (Le et al. Reference Le, Egedal, Daughton, Roytershteyn, Karimabadi and Forest2015).

The kinetic investigation of current sheet stability with the effects of particle effective collisions requires the consideration of the Vlasov–Maxwell equations with a collision integral (e.g. Zelenyi & Artemyev Reference Zelenyi and Artemyev2013, and references therein). Due to the complicated form of the full collision integral (Pitaevskii & Lifshitz Reference Pitaevskii and Lifshitz1981), the practical approach is reduced to the inclusion of an approximate form of this integral into a kinetic equation. For example, one of the most popular forms of the collision operator $Y[f]$ acting on the velocity distribution $f=f_{0}+{\it\delta}f$ has been proposed by Bhatnagar, Gross & Krook (Reference Bhatnagar, Gross and Krook1954):

(1.1) $$\begin{eqnarray}Y[f]=-{\it\nu}\left({\it\delta}f-\left(\frac{n}{n_{0}}+\frac{m}{T}\left\langle \boldsymbol{v}\right\rangle \boldsymbol{v}\right)f_{0}\right),\end{eqnarray}$$

where ${\it\nu}$ is Coulomb collision frequency (later treated more generally as a frequency of effective collisions due to interaction of electrons with different wave modes), $m$ and $T$ are the mass and temperature of the particles, $f_{0}$ and $n_{0}$ are the unperturbed velocity distribution and plasma density and

(1.2) $$\begin{eqnarray}\left.\begin{array}{@{}c@{}}\displaystyle \left\langle \boldsymbol{v}\right\rangle =n_{0}^{-1}\int \boldsymbol{v}{\it\delta}f\,\text{d}\boldsymbol{v}\\ \displaystyle n=\int {\it\delta}f\,\text{d}\boldsymbol{v}.\end{array}\right\}\end{eqnarray}$$

We also note that $\int \!\boldsymbol{v}f_{0}\,\text{d}\boldsymbol{v}=0$ and $n_{0}=\int \!f_{0}\,\text{d}\boldsymbol{v}$ .

The integral (1.1) describes the relaxation of the velocity distributions to the initial state $f_{0}$ . This integral does not include any derivatives of the perturbation ${\it\delta}f$ and thus does not take into account any information about the internal fine structure of ${\it\delta}f$ . In this paper, we show how the consideration of exact form of the collision integral can modify the effective frequency ${\it\nu}$ with particle gyration in the background magnetic field produces significant modulations in the distribution function ${\it\delta}f$ .

2. Collisions in a weakly magnetized plasma

To consider the effect of a finite electron gyroradius on effective collisions below we use the full expression of the Landau form of the collision integral $Y[f]=-m^{-1}(\boldsymbol{{\rm\nabla}}\boldsymbol{\cdot }\boldsymbol{s})$ where the density of particle flux $\boldsymbol{s}$ has the following components (Pitaevskii & Lifshitz Reference Pitaevskii and Lifshitz1981):

(2.1) $$\begin{eqnarray}\displaystyle s_{{\it\alpha}} & = & \displaystyle \frac{{\rm\pi}e^{2}{\it\Lambda}}{m}\mathop{\sum }_{{\it\beta}}\int \left(\frac{\partial f(\boldsymbol{v}^{\prime })}{\partial v_{{\it\beta}}^{\prime }}+\frac{mv_{{\it\beta}}^{\prime }}{T}f(\boldsymbol{v}^{\prime })\right)f_{0}(\boldsymbol{v})\frac{w^{2}{\it\delta}_{{\it\beta}{\it\alpha}}-w_{{\it\alpha}}w_{{\it\beta}}}{w^{2}}\,\text{d}\boldsymbol{v}^{\prime }\nonumber\\ \displaystyle & & \displaystyle -\,\frac{{\rm\pi}e^{2}{\it\Lambda}}{m}\mathop{\sum }_{{\it\beta}}\int \left(\frac{\partial f(\boldsymbol{v})}{\partial v_{{\it\beta}}}+\frac{mv_{{\it\beta}}}{T}f(\boldsymbol{v})\right)f_{0}(\boldsymbol{v}^{\prime })\frac{w^{2}{\it\delta}_{{\it\beta}{\it\alpha}}-w_{{\it\alpha}}w_{{\it\beta}}}{w^{2}}\,\text{d}\boldsymbol{v}^{\prime },\end{eqnarray}$$

where $\boldsymbol{w}=\boldsymbol{v}-\boldsymbol{v}^{\prime }$ , ${\it\alpha},{\it\beta}=x,y,z$ , ${\it\Lambda}=\ln ({\it\lambda}_{D}q^{2}/T)$ and ${\it\lambda}_{D}$ is the Debye length.

Considering the perturbation ${\it\delta}f=f(\boldsymbol{v})\exp (\text{i}{\it\Phi})$ of the initial distribution function $f_{0}(\boldsymbol{v})$ , the phase of the perturbation is ${\it\Phi}=\boldsymbol{k}\boldsymbol{\cdot }\boldsymbol{r}-{\it\omega}t$ . The corresponding perturbations to the electromagnetic field are ${\it\delta}\boldsymbol{B}=\boldsymbol{B}\exp (\text{i}{\it\Phi})$ , ${\it\delta}\boldsymbol{E}=\boldsymbol{E}\exp (\text{i}{\it\Phi})$ , while the background magnetic field is $\boldsymbol{B}_{0}=B_{0}\boldsymbol{e}_{z}$ . The wave vector of perturbations $\boldsymbol{k}$ is assumed to lie in the $(x,z)$ plane. In this case, the perturbation of the Valsov equation takes the form

(2.2) $$\begin{eqnarray}\frac{\partial {\it\delta}f}{\partial t}+\boldsymbol{v}\frac{\partial {\it\delta}f}{\partial \boldsymbol{r}}+\frac{q}{mc}[\boldsymbol{v}\times \boldsymbol{B}_{0}]\frac{\partial {\it\delta}f}{\partial \boldsymbol{v}}+Y[{\it\delta}f]=-\frac{q}{m}\left({\it\delta}\boldsymbol{E}+\frac{1}{c}[\boldsymbol{v}\times {\it\delta}\boldsymbol{B}]\right)\frac{\partial f_{0}}{\partial \boldsymbol{v}}.\end{eqnarray}$$

We introduce the force $\boldsymbol{F}=(q/m)(\boldsymbol{E}+c^{-1}[\boldsymbol{v}\times \boldsymbol{B}])$ , cylindrical velocity coordinates $v_{x}=v_{\bot }\cos {\it\theta}$ , $v_{y}=v_{\bot }\sin {\it\theta}$ and rewrite (2.2) as

(2.3) $$\begin{eqnarray}-\text{i}{\it\omega}{\it\delta}f+\text{i}k_{z}v_{z}{\it\delta}f+\text{i}k_{x}v_{\bot }\cos {\it\theta}{\it\delta}f-\frac{qB_{0}}{mc}\frac{\partial {\it\delta}f}{\partial {\it\theta}}+Y[{\it\delta}f]=-\frac{q}{m}\boldsymbol{F}\text{e}^{\text{i}{\it\Phi}}\frac{\partial f_{0}}{\partial \boldsymbol{v}},\end{eqnarray}$$

where $[\boldsymbol{v}\times \boldsymbol{B}_{0}](\partial /\partial \boldsymbol{v})=-B_{0}(\partial /\partial {\it\theta})$ . Then we divide (2.3) by $\exp (\text{i}{\it\Phi})$ and introduce ${\it\Omega}_{0}=qB_{0}/mc$ , ${\it\lambda}={\it\omega}-k_{z}v_{z}$ :

(2.4) $$\begin{eqnarray}-\text{i}({\it\lambda}-k_{x}v_{\bot }\cos {\it\theta})f-{\it\Omega}_{0}\frac{\partial f}{\partial {\it\theta}}+Y[f]=-\boldsymbol{F}\frac{\partial f_{0}}{\partial \boldsymbol{v}}.\end{eqnarray}$$

We introduce the function $W=({\it\lambda}{\it\theta}-k_{x}v_{\bot }\cos {\it\theta})/{\it\Omega}_{0}$ and function $g(\boldsymbol{v})=f(\boldsymbol{v})\exp (\text{i}W)$ . Thus, (2.4) can be rewritten as

(2.5) $$\begin{eqnarray}\frac{\partial g}{\partial {\it\theta}}+\text{e}^{\text{i}W}Y[g\text{e}^{-\text{i}W}]=-\boldsymbol{F}\frac{\partial f_{0}}{\partial \boldsymbol{v}}\text{e}^{\text{i}W}.\end{eqnarray}$$

To derive the dispersion relation for perturbations, one should substitute into (2.5) the Maxwell equations for electromagnetic field perturbations ( ${\sim}\boldsymbol{F}$ ) expressed through function $g$ (see, e.g. Zelenyi & Artemyev Reference Zelenyi and Artemyev2013). In this case, the final wave frequency/growth rate would depend on collision frequency (e.g. for the simplified collision integral (1.1) the operator $Y$ is proportional to collision frequency  ${\it\nu}$ ). However, we would like to estimate the effect of a finite electron gyroradius on collision frequency ${\it\nu}$ . Thus, we compare the term ${\sim}Y$ for the two systems: when simplified (1.1) can be used and $Y\sim {\it\nu}$ and when the full collision integral $Y[f]=-m^{-1}(\boldsymbol{{\rm\nabla}}\boldsymbol{\cdot }\boldsymbol{s})$ with (2.1) should be taken into account. For this reason, we carefully consider the second term in (2.5) and estimate the main part of this term

(2.6) $$\begin{eqnarray}\int ^{{\it\theta}}\text{e}^{\text{i}W}Y[g\text{e}^{-\text{i}W}]\,\text{d}{\it\theta}.\end{eqnarray}$$

In the limit $k_{x}\sqrt{2T/m}/{\it\Omega}_{0}\gg 1$ , function $g\exp (-\text{i}W)$ contains the fast oscillating term ${\sim}\exp (\text{i}k_{x}v_{\bot }\sin {\it\theta}/{\it\Omega}_{0})=\exp (\text{i}k_{x}v_{y}/{\it\Omega}_{0})$ . Thus, in (2.1) we should keep the main terms corresponding to the derivative $\partial f/\partial v_{y}$ . The first term in (2.1) contains the integral $\int (\partial f/\partial v_{y}^{\prime })\,\text{d}v_{y}^{\prime }\sim f$ . Therefore, the second term with $\partial f/\partial v_{y}$ is more important and we can write:

(2.7) $$\begin{eqnarray}\displaystyle & \displaystyle s_{{\it\alpha}}\approx -\frac{{\rm\pi}e^{2}{\it\Lambda}}{m}\left(\frac{\partial f(\boldsymbol{v})}{\partial v_{y}}+\frac{mv_{y}}{T}f(\boldsymbol{v})\right)A_{{\it\alpha}}(\boldsymbol{v}) & \displaystyle\end{eqnarray}$$
(2.8) $$\begin{eqnarray}\displaystyle & \displaystyle A_{{\it\alpha}}(\boldsymbol{v})=\int f_{0}(\boldsymbol{v}^{\prime })\frac{w^{2}{\it\delta}_{y{\it\alpha}}-w_{y}w_{{\it\beta}}}{w^{2}}\,\text{d}\boldsymbol{v}^{\prime }. & \displaystyle\end{eqnarray}$$

The main part of $s_{{\it\alpha}}$ from (2.8) is

(2.9) $$\begin{eqnarray}s_{{\it\alpha}}\approx -\text{i}\frac{k_{x}}{{\it\Omega}_{0}}\frac{{\rm\pi}e^{2}{\it\Lambda}}{m}g(\boldsymbol{v})\text{e}^{-\text{i}W}A_{{\it\alpha}}(\boldsymbol{v}).\end{eqnarray}$$

Thus, the integrand of (2.6) takes the form

(2.10) $$\begin{eqnarray}\text{e}^{\text{i}W}Y[g\text{e}^{-\text{i}W}]=-\frac{k_{x}^{2}}{{\it\Omega}_{0}^{2}}\frac{{\rm\pi}e^{2}{\it\Lambda}}{m^{2}}g(\boldsymbol{v})A_{{\it\alpha}}(\boldsymbol{v}).\end{eqnarray}$$

In dimensionless form (2.10) can be written as

(2.11) $$\begin{eqnarray}\text{e}^{\text{i}W}Y[g\text{e}^{-\text{i}W}]=-\frac{2k_{x}^{2}T}{{\it\Omega}_{0}^{2}m}{\tilde{Y}}[f]\text{e}^{\text{i}W},\end{eqnarray}$$

where ${\tilde{Y}}$ is the initial collision integral (2.1) without derivatives over fast oscillation terms. Equation (2.11) shows that the effect of a finite electron gyroradius provides the multiplication factor $2k_{x}^{2}T/{\it\Omega}_{0}^{2}m$ . This factor is omitted in the simplified form of the collision integral (1.1). Therefore, if we operate with the collision frequency ${\it\nu}$ from (1.1) then this expression for collision frequency ${\it\nu}$ (Coulomb or effective) should be multiplied by the term $2k_{x}^{2}T/{\it\Omega}_{0}^{2}m\gg 1$ . Of course, for a complete investigation of system stability one should consider the full collision integral (2.1), but for many applications the simplified approach with collision frequency ${\it\nu}$ can be applied with the corresponding correction ${\sim}2k_{x}^{2}T/{\it\Omega}_{0}^{2}m$ . Finally, we come to the following expression for the effective collision frequency in a system with weakly magnetized electrons:

(2.12) $$\begin{eqnarray}{\it\nu}_{eff}\approx \left\{\begin{array}{@{}ll@{}}{\it\nu}(k_{x}{\it\rho}_{e})^{2}\quad & k_{x}{\it\rho}_{e}>1\\ {\it\nu}\quad & k_{x}{\it\rho}_{e}\leqslant 1,\end{array}\right.\end{eqnarray}$$

where ${\it\rho}_{e}=\sqrt{2T/m}/{\it\Omega}_{0}$ is the electron gyroradius. Note, that (2.12) provides only a simplified estimate of the effect of electron finite gyroradius on system stability. However, this simplified expression gives us the opportunity to estimate this effect for realistic system parameters. In the next section we use the modified collision frequency (2.12) to estimate the effect of effective conductivity in the regions with weak magnetic field, in particular in the reconnection regions in near-Earth plasma systems.

3. Estimates of $k_{x}{\it\rho}_{e}$ parameter for space plasma reconnection systems

Let us consider the classical 2-D X-line with magnetic field configuration $\boldsymbol{B}=B_{0}(z/L_{z})\boldsymbol{e}_{x}+B_{z}(x/L_{x})\boldsymbol{e}_{z}$ (see scheme in figure 1). The reversal of the $B_{x}$ magnetic field component in the neutral plane $z=0$ generates so-called neutral region $|z|<\sqrt{{\it\rho}_{0}L_{z}}$ with ${\it\rho}_{0}=\sqrt{2Tm}c/eB_{0}$ where particles are not affected by the $B_{x}$ magnetic field and can move practically freely along the $y$ -axis (Dobrowolny Reference Dobrowolny1968). The same region appears around the $x=0$ plane due to $B_{z}$ reversal: $|x|<\sqrt{{\it\rho}_{z}L_{x}}$ with ${\it\rho}_{z}=\sqrt{2Tm}c/eB_{z}$ . Particles, moving within this region are affected by the effective (averaged gyroradius) magnetic field with amplitude $\tilde{B}_{z}=B_{z}\sqrt{{\it\rho}_{z}L_{x}}/L_{x}=B_{z}\sqrt{{\it\rho}_{z}/L_{x}}$ , while the effective particle gyroradius is equal to the spatial scale ${\it\rho}_{e}=\sqrt{2Tm}c/e\tilde{B}_{z}=\sqrt{{\it\rho}_{z}L_{x}}$ . If we assume that the X-line is generated by the tearing instability with a wavelength equal to $L_{x}=2{\rm\pi}/k_{x}$ (see the scheme in figure 1), then the factor from (2.12) can be written as

(3.1) $$\begin{eqnarray}X=(k_{x}{\it\rho}_{e})^{2}={\it\rho}_{z}L_{x}\left(\frac{2{\rm\pi}}{L_{x}}\right)^{2}=4{\rm\pi}^{2}\frac{{\it\rho}_{z}}{L_{x}}.\end{eqnarray}$$

Figure 2 shows values of $X$ as a function of electron temperature $T$ , $L_{x}$ and $B_{z}$ . For the initial stage of the tearing instability (when $B_{z}$ is still very small) with a wavelength $L_{x}$ which is not too long the factor $X$ is larger than one. Thus, the effect of a finite electron gyroradius can amplify energy dissipation due to effective collisions.

Figure 1. Schematic view of the X-line region.

Figure 2. Factor $X$ as a function of $L_{x}$ for two values of $B_{z}$ and four values of electron temperature $T$ .

To estimate the effect of the factor $X$ on magnetic reconnection in real plasma systems we consider the effective collisions induced by two types of turbulence: lower-hybrid drift (LHD) turbulence (Huba et al. Reference Huba, Gladd and Papadopoulos1977) and kinetic Alfvén wave (KAW) turbulence (Chaston et al. Reference Chaston, Johnson, Wilber, Acuna, Goldstein and Reme2009).

Wavelengths of the electromagnetic LHD mode ( ${\sim}{\it\rho}_{e}\sqrt{m_{i}/m_{e}}$ (Daughton Reference Daughton2003)) and KAW ( ${<}0.3{\it\rho}_{e}(m_{i}/me)$ (Voitenko Reference Voitenko1998; Chen et al. Reference Chen, Wu, Zhao, Tang and Huang2014)) are small enough to consider these waves to be small-scale magnetic field perturbations for the large-scale reconnection process. On the other hand, these modes effectively interact both with ions (through the Cherenkov resonance $\boldsymbol{k}\boldsymbol{v}={\it\omega}$ (Karney Reference Karney1978; Karimabadi et al. Reference Karimabadi, Akimoto, Omidi and Menyuk1990; Chaston et al. Reference Chaston, Bonnell, Wygant, Mozer, Bale, Kersten, Breneman, Kletzing, Kurth and Hospodarsky2014)) and electrons (through the Landau resonance $k_{\Vert }v_{\Vert }={\it\omega}$ (Hasegawa Reference Hasegawa1976; Cairns & McMillan Reference Cairns and McMillan2005)). Both modes are widely observed in the vicinity of the reconnection region where their contributions to the electromagnetic field turbulence are the strongest among wave modes (Eastwood et al. Reference Eastwood, Phan, Bale and Tjulin2009; Fujimoto, Shinohara & Kojima Reference Fujimoto, Shinohara and Kojima2011; Huang et al. Reference Huang, Zhou, Sahraoui, Vaivads, Deng, André, He, Fu, Li and Yuan2012). Thus, LHD and KAW can the provide an exchange of energy between the ions and electrons, supporting the anomalous (effective) conductivity. The general form of the frequency of effective collisions ${\it\nu}$ is provided by the quasi-linear equation (Galeev & Sagdeev Reference Galeev, Sagdeev and Leontovich1979):

(3.2) $$\begin{eqnarray}{\it\nu}=\frac{1}{mn_{0}v_{D}}\int W_{k}\frac{{\it\Gamma}(k)}{{\it\omega}(k)}k\frac{\text{d}^{3}k}{(2{\rm\pi})^{3}},\end{eqnarray}$$

where $v_{D}$ is particle drift velocity (i.e. $en_{0}v_{D}$ is a current density), ${\it\Gamma}(k)$ is a wave growth/damping rate, ${\it\omega}(k)$ is a wave frequency and $W_{k}$ is a wave energy density. For current sheet configurations the maximum value of ${\it\nu}_{LHD}$ was derived by Huba, Drake & Gladd (Reference Huba, Drake and Gladd1980):

(3.3) $$\begin{eqnarray}{\it\nu}_{LHD}\approx \frac{{\it\omega}_{pe}^{2}}{{\it\omega}_{LH}}\frac{E_{y}^{2}}{8{\rm\pi}n_{0}T_{e}},\end{eqnarray}$$

where ${\it\omega}_{pe}$ is the plasma frequency, ${\it\omega}_{LH}\approx {\it\Omega}_{e}\sqrt{m_{e}/m_{i}}$ and ${\it\Omega}_{e}=eB_{z}/m_{e}c$ is the electron gyrofrequency, $T_{e}$ is electron temperature and $E_{y}$ corresponds to the amplitude of wave electric field oscillations.

For KAW the maximum growth rate corresponds to Landau excitation ${\it\Gamma}\approx {\it\omega}/kv_{D}$ (Hasegawa & Mima Reference Hasegawa and Mima1978) where the wave frequency is ${\it\omega}_{KAW}=k_{\Vert }v_{A}\sqrt{1+(k_{\bot }{\it\rho}_{i})^{2}}$ with ion gyroradius ${\it\rho}_{i}$ and Alfvén velocity $v_{A}$ (Hasegawa Reference Hasegawa1976). The corresponding collision frequency can be estimated as

(3.4) $$\begin{eqnarray}{\it\nu}_{KAW}\approx {\it\omega}_{KAW}{\it\Gamma}\frac{{\it\omega}_{pe}^{2}}{k_{\Vert }^{2}v_{Te}^{2}}\frac{E_{y}^{2}}{4{\rm\pi}n_{0}m_{e}v_{D}^{2}}\approx \frac{{\it\omega}_{pe}^{2}}{{\it\omega}_{KAW}}\frac{v_{A}^{3}}{v_{D}^{2}v_{Ti}}\frac{E_{y}^{2}}{8{\rm\pi}n_{0}T_{e}}\approx \frac{{\it\omega}_{pe}^{2}}{{\it\omega}_{KAW}}\frac{E_{y}^{2}}{8{\rm\pi}n_{0}T_{e}},\end{eqnarray}$$

where we take into account that $k_{\bot }{\it\rho}_{i}\sim 1$ , $v_{D}\sim v_{Ti}\sim v_{A}$ .

Following Coroniti (Reference Coroniti1985), one can determine the critical amplitude of electric field energy $E_{y}^{2}$ necessary to organize the magnetic field dissipation within the domain with spatial scale ${\sim}L_{z}\sim {\it\rho}_{i}$ :

(3.5) $$\begin{eqnarray}\displaystyle & \displaystyle \frac{(E_{y}^{2})_{LHD}}{8{\rm\pi}n_{0}T_{e}}=\frac{M_{A}}{X}\sqrt{\frac{m_{i}}{m_{e}}}\frac{v_{A}v_{Ti}}{c^{2}}\approx \frac{M_{A}}{X}\sqrt{\frac{m_{i}}{m_{e}}}\left(\frac{v_{A}}{c}\right)^{2} & \displaystyle\end{eqnarray}$$
(3.6) $$\begin{eqnarray}\displaystyle & \displaystyle \frac{(E_{y}^{2})_{KAW}}{8{\rm\pi}n_{0}T_{e}}=\frac{M_{A}}{X}\frac{{\it\omega}_{KAW}}{{\it\Omega}_{i}}\left(\frac{v_{A}}{c}\right)^{2}\approx \frac{M_{A}}{X}\left(\frac{v_{A}}{c}\right)^{2}, & \displaystyle\end{eqnarray}$$

where $M_{A}$ is an Alfvén–Mach number for a particle flowing to the reconnection region. Generally $M_{A}\sim 0.1$ for magnetospheric physics, while $(v_{A}/c)^{2}\approx 10^{-5}$ for the Earth’s magnetotail. Equation (3.6) shows that one can decrease the estimate for the critical amplitude of wave electric field proportionally to ${\sim}1/\sqrt{X}\sim 1/3{-}1/10$ (see figure 2) after taking into account the Pitaevskii effect. For magnetotail current sheets the typical amplitude of LHD waves can reach $10~\text{mV}~\text{m}^{-1}$ (Eastwood et al. Reference Eastwood, Phan, Bale and Tjulin2009; Fujimoto et al. Reference Fujimoto, Shinohara and Kojima2011; Norgren et al. Reference Norgren, Vaivads, Khotyaintsev and André2012), while the estimate of the amplitude from (3.6) is $\sqrt{(E_{y}^{2})_{LHD}}\sim 30/\sqrt{X}~\text{mV}~\text{m}^{-1}$ (Coroniti Reference Coroniti1985). Thus, a factor $X\sim 10$ can help to produce the necessary magnetic energy dissipation due to effective collisions. Amplitudes of KAW electric field are often weaker ${\sim}1~\text{mV}~\text{m}^{-1}$ (Chaston et al. Reference Chaston, Bonnell, Clausen and Angelopoulos2012). The critical amplitude for KAW is $\sqrt{(E_{y}^{2})_{KAW}}$ is of the order of ${\sim}1/\sqrt{X}~\text{mV}~\text{m}^{-1}$ . The Pitaevskii effect of reducing the estimate (3.6) 2–3 times demonstrates an important role of KAW in running magnetic reconnection. Therefore, we conclude that in realistic magnetotail conditions for both cases (LHDI and KAW) the effect of finite electron gyroradius can increase effective collisions and help to overturn Coroniti (Reference Coroniti1985) objections.

4. Discussion and conclusions

The additional effect of the enhancement of effective collisions for $k{\it\rho}_{e}>1$ perturbations was taken into account in the analysis of the electron resistive tearing mode (Zelenyi & Taktakishvili Reference Zelenyi and Taktakishvili1981). This mode is excited by ${\it\nu}\neq 0$ with the growth rate ${\it\gamma}={\it\nu}\text{X}({\it\gamma}_{0}/k_{x}\sqrt{2T_{e}m_{e}})$ where ${\it\gamma}_{0}/k_{x}\sqrt{2T_{e}m_{e}}=2{\rm\pi}^{-1/2}({\it\rho}_{e}/L_{z})^{3/2}(1-k_{x}^{2}L_{z}^{2})/k_{x}L_{z}$ is the growth rate of the electron tearing mode (Laval et al. Reference Laval, Pellat and Vuillemin1966; Galeev & Sudan Reference Galeev and Sudan1985). Therefore, for $X$ factor high enough, the growth rate of the resistive tearing mode increases. This mode is produced by effective collisions and cannot be stabilized by electron magnetization (in contrast to the classical electron tearing mode (Schindler Reference Schindler1974; Galeev & Zelenyi Reference Galeev and Zelenyi1976)). Moreover, for a tearing mode with $k{\it\rho}_{e}>1$ the WKB approximation in the investigation of current sheet stability can be safely applied (Lembege & Pellat Reference Lembege and Pellat1982) supporting the validity of the expressions for the growth rates ${\it\gamma}$ and ${\it\gamma}_{0}$ derived for the case where the wavelength is not too long.

Effective collisions are often considered in numerical models of magnetic reconnection as a trigger for the reconnection process (Daughton, Lapenta & Ricci Reference Daughton, Lapenta and Ricci2004; Daughton et al. Reference Daughton, Roytershteyn, Karimabadi, Yin, Albright, Bergen and Bowers2011; Karimabadi et al. Reference Karimabadi, Roytershteyn, Daughton and Liu2013). In this case, estimates of the reconnection rate can be based on the classical theory of an effective conductivity (Galeev & Sudan Reference Galeev and Sudan1985; Yoon & Lui Reference Yoon and Lui2006) with a proper estimate for the effective collision rate. Thus, the effect of fine structured velocity distributions (when ${\it\rho}k_{x}>1$ ) can be very important. The same effect can be provided by numerical resistivity which supports the slow growth of magnetic islands when the tearing instability is saturated by nonlinear effects (Lipatov & Zelenyi Reference Lipatov and Zelenyi1982).

Figure 2 shows that the most pronounced effect of a finite electron gyroradius corresponds to small-scale reconnection (i.e. small $L_{x}$ ) where instead of one large-scale X-line we deal with a chain of small-scale magnetic islands. Indeed, strong anisotropy of electrons accelerated in the primary reconnection region generates high-amplitude curvature currents (Egedal, Le & Daughton Reference Egedal, Le and Daughton2013; Artemyev et al. Reference Artemyev, Petrukovich, Nakamura and Zelenyi2015) responsible for the formation of very thin current sheets with vanishing $B_{z}$ magnetic field (Nakamura et al. Reference Nakamura, Baumjohann, Runov and Asano2006; Artemyev et al. Reference Artemyev, Petrukovich, Frank, Nakamura and Zelenyi2013; Le et al. Reference Le, Egedal, Ng, Karimabadi, Scudder, Roytershteyn, Daughton and Liu2014). Instability of such thin current sheets results in secondary reconnection in the outflow region of the primary large-scale X-line (so-called plasmoid instability) with the corresponding birth of a multitude of small-scale X- and O-magnetic points (Daughton et al. Reference Daughton, Roytershteyn, Karimabadi, Yin, Albright, Bergen and Bowers2011; Huang, Bhattacharjee & Forbes Reference Huang, Bhattacharjee and Forbes2013). Thus, the secondary reconnection of such thin current sheets can occur within the region of $k_{x}{\it\rho}_{e}>1$ . To illustrate this scenario we show two events of Cluster spacecraft observations of magnetic reconnection in the Earth’s magnetotail (see figure 3). The characteristic $B_{z}$ reversal and plasma flows $v_{x}$ indicate that Cluster is in close vicinity to the X-line. We calculate the electron gyroradius ${\it\rho}_{e}$ using local measurements of the $B_{z}$ and $B_{y}$ magnetic field ( $B_{x}$ is assumed to be zero). Around the X-line ${\it\rho}_{e}$ reaches ${\sim}1000~\text{km}$ , thus $X$ becomes larger than one for $L_{x}<6R_{E}$ . For example, the secondary reconnection with a wavelength of approximately ${\sim}2R_{E}$ corresponds to an increase of effective collision frequency by a factor $X\sim 3$ . We also note that for both events shown in figure 3, current sheet thicknesses were approximately ${\sim}800~\text{km}$ and ${\sim}600~\text{km}$ (see estimates of current density amplitudes in Artemyev et al. Reference Artemyev, Petrukovich, Nakamura and Zelenyi2015). Thus, for $L_{x}\sim 2R_{E}$ the ratio $L_{x}/L_{z}\sim 18$ and the corresponding current sheets are very prolonged and stretched.

Figure 3. Two cases of Cluster spacecraft crossing a reconnection region (see the characteristic reversal of $B_{z}$ ). Both cases are taken from Artemyev et al. (Reference Artemyev, Petrukovich, Nakamura and Zelenyi2015).

To conclude, we consider the effects of a finite electron gyroradius for current sheet stability in the case when effective collisions are present in the system. Following Pitaevskii (Reference Pitaevskii1963) we demonstrate that a fast electron gyrorotation results in a fine structure of the perturbations of the electron distribution function. As a result, the full expression for the collision integral gives a multiplication factor $\text{X}\sim (k_{x}{\it\rho}_{e})^{2}$ for collision frequency (where $k_{x}$ is a transverse wavenumber). For realistic conditions in the Earth’s magnetotail this factor can often be larger than one and thus can increase the effect of particle scattering by electromagnetic turbulence. We show that effect of a finite electron gyroradius should be particularly strong for the secondary reconnection of very thin current sheets formed in the outflow region of the primary X-line.

Acknowledgement

Work of A.V.A. was supported by Russian Scientific Foundation (project no. 14-12-00824).

References

Angelopoulos, V., McFadden, J. P., Larson, D., Carlson, C. W., Mende, S. B., Frey, H., Phan, T., Sibeck, D. G., Glassmeier, K.-H., Auster, U. et al. 2008 Tail reconnection triggering substorm onset. Science 321, 931935.Google Scholar
Artemyev, A. V., Petrukovich, A. A., Frank, A. G., Nakamura, R. & Zelenyi, L. M. 2013 Intense current sheets in the magnetotail: peculiarities of electron physics. J. Geophys. Res. 118, 27892799.Google Scholar
Artemyev, A. V., Petrukovich, A. A., Nakamura, R. & Zelenyi, L. M. 2015 Statistics of intense dawn-dusk currents in the Earth’s magnetotail. J. Geophys. Res. 120, 38043820.Google Scholar
Baumjohann, W. 2002 Modes of convection in the magnetotail. Phys. Plasmas 9, 36653667.Google Scholar
Bhatnagar, P. L., Gross, E. P. & Krook, M. 1954 A model for collision processes in gases. I. Small amplitude processes in charged and neutral one-component systems. Phys. Rev. 94, 511525.CrossRefGoogle Scholar
Büchner, J. & Zelenyi, L. M. 1987 Chaotization of the electron motion as the cause of an internal magnetotail instability and substorm onset. J. Geophys. Res. 92, 1345613466.Google Scholar
Cairns, I. H. & McMillan, B. F. 2005 Electron acceleration by lower hybrid waves in magnetic reconnection regions. Phys. Plasmas 12, 102110.CrossRefGoogle Scholar
Chaston, C. C., Bonnell, J. W., Clausen, L. & Angelopoulos, V. 2012 Energy transport by kinetic-scale electromagnetic waves in fast plasma sheet flows. J. Geophys. Res. 117, 9202.Google Scholar
Chaston, C. C., Bonnell, J. W., Wygant, J. R., Mozer, F., Bale, S. D., Kersten, K., Breneman, A. W., Kletzing, C. A., Kurth, W. S., Hospodarsky, G. B. et al. 2014 Observations of kinetic scale field line resonances. Geophys. Res. Lett. 41, 209215.Google Scholar
Chaston, C. C., Johnson, J. R., Wilber, M., Acuna, M., Goldstein, M. L. & Reme, H. 2009 Kinetic Alfvén wave turbulence and transport through a reconnection diffusion region. Phys. Rev. Lett. 102, 015001.Google Scholar
Chen, L., Wu, D. J., Zhao, G. Q., Tang, J. F. & Huang, J. 2014 Excitation of kinetic Alfvén waves by fast electron beams. Astrophys. J. 793, 13.Google Scholar
Coroniti, F. V. 1980 On the tearing mode in quasi-neutral sheets. J. Geophys. Res. 85, 67196728.Google Scholar
Coroniti, F. V. 1985 Space plasma turbulent dissipation – Reality or myth? Space Sci. Rev. 42, 399410.Google Scholar
Daughton, W. 2003 Electromagnetic properties of the lower-hybrid drift instability in a thin current sheet. Phys. Plasmas 10, 31033119.Google Scholar
Daughton, W., Lapenta, G. & Ricci, P. 2004 Nonlinear evolution of the lower-hybrid drift instability in a current sheet. Phys. Rev. Lett. 93, 105004.Google Scholar
Daughton, W., Roytershteyn, V., Karimabadi, H., Yin, L., Albright, B. J., Bergen, B. & Bowers, K. J. 2011 Role of electron physics in the development of turbulent magnetic reconnection in collisionless plasmas. Nat. Phys. 7, 539542.CrossRefGoogle Scholar
Dobrowolny, M. 1968 Instability of a neutral sheet. Nuovo Cimento B 55, 427442.Google Scholar
Eastwood, J. P., Phan, T. D., Bale, S. D. & Tjulin, A. 2009 Observations of turbulence generated by magnetic reconnection. Phys. Rev. Lett. 102, 035001.Google Scholar
Egedal, J., Le, A. & Daughton, W. 2013 A review of pressure anisotropy caused by electron trapping in collisionless plasma, and its implications for magnetic reconnection. Phys. Plasmas 20, 061201.Google Scholar
Fujimoto, M., Shinohara, I. & Kojima, H. 2011 Reconnection and waves: a review with a perspective. Space Sci. Rev. 160, 123143.Google Scholar
Galeev, A. A. & Sagdeev, R. Z. 1979 Nonlinear plasma theory. In Reviews of Plasma Physics (ed. Leontovich, A. M. A.), vol. 7, p. 1. Springer.Google Scholar
Galeev, A. A. & Sudan, R. N. 1985 Basic plasma physics II. In Handbook of Plasma Physics, vol. 2. Elsevier.Google Scholar
Galeev, A. A. & Zelenyi, L. M. 1976 Tearing instability in plasma configurations. Sov. J. Expl Theor. Phys. 43, 1113.Google Scholar
Greco, A., Valentini, F., Servidio, S. & Matthaeus, W. H. 2012 Inhomogeneous kinetic effects related to intermittent magnetic discontinuities. Phys. Rev. E 86, 066405.Google Scholar
Hasegawa, A. 1976 Particle acceleration by MHD surface wave and formation of aurora. J. Geophys. Res. 81, 50835090.Google Scholar
Hasegawa, A. & Mima, K. 1978 Anomalous transport produced by kinetic Alfvén wave turbulence. J. Geophys. Res. 83, 11171123.Google Scholar
Huang, S. Y., Zhou, M., Sahraoui, F., Vaivads, A., Deng, X. H., André, M., He, J. S., Fu, H. S., Li, H. M., Yuan, Z. G. et al. 2012 Observations of turbulence within reconnection jet in the presence of guide field. Geophys. Res. Lett. 39, 11104.CrossRefGoogle Scholar
Huang, Y.-M., Bhattacharjee, A. & Forbes, T. G. 2013 Magnetic reconnection mediated by hyper-resistive plasmoid instability. Phys. Plasmas 20, 082131.Google Scholar
Huba, J. D., Drake, J. F. & Gladd, N. T. 1980 Lower-hibrid-drift instability in field reversed plasmas. Phys. Fluids 23, 552561.Google Scholar
Huba, J. D., Gladd, N. T. & Papadopoulos, K. 1977 The lower-hybrid-drift instability as a source of anomalous resistivity for magnetic field line reconnection. Geophys. Res. Lett. 4, 125126.Google Scholar
Karimabadi, H., Akimoto, K., Omidi, N. & Menyuk, C. R. 1990 Particle acceleration by a wave in a strong magnetic field – Regular and stochastic motion. Phys. Fluids B 2, 606628.Google Scholar
Karimabadi, H., Roytershteyn, V., Daughton, W. & Liu, Y.-H. 2013 Recent evolution in the theory of magnetic reconnection and its connection with turbulence. Space Sci. Rev. 178, 307323.Google Scholar
Karney, C. F. F. 1978 Stochastic ion heating by a lower hybrid wave. Phys. Fluids 21, 15841599.CrossRefGoogle Scholar
Laval, G., Pellat, R. & Vuillemin, M. 1966 Instabilités électromagnétiques des plasmas sans collisions (CN- $21/71$ ). In Plasma Physics and Controlled Nuclear Fusion Research, vol. II, pp. 259277. International Atomic Energy Agency.Google Scholar
Le, A., Egedal, J., Daughton, W., Roytershteyn, V., Karimabadi, H. & Forest, C. 2015 Transition in electron physics of magnetic reconnection in weakly collisional plasma. J. Plasma Phys. 81, 30108.Google Scholar
Le, A., Egedal, J., Ng, J., Karimabadi, H., Scudder, J., Roytershteyn, V., Daughton, W. & Liu, Y.-H. 2014 Current sheets and pressure anisotropy in the reconnection exhaust. Phys. Plasmas 21, 012103.CrossRefGoogle Scholar
Lembege, B. & Pellat, R. 1982 Stability of a thick two-dimensional quasineutral sheet. Phys. Fluids 25, 19952004.Google Scholar
Lipatov, A. S & Zelenyi, L. M. 1982 The study of magnetic islands dynamics. Plasma Phys. 24, 10821089.Google Scholar
Nagai, T., Shinohara, I., Fujimoto, M., Matsuoka, A., Saito, Y. & Mukai, T. 2011 Construction of magnetic reconnection in the near-Earth magnetotail with Geotail. J. Geophys. Res. 116, 4222.Google Scholar
Nakamura, R., Baumjohann, W., Runov, A. & Asano, Y. 2006 Thin current sheets in the magnetotail observed by cluster. Space Sci. Rev. 122, 2938.Google Scholar
Norgren, C., Vaivads, A., Khotyaintsev, Y. V. & André, M. 2012 Lower hybrid drift waves: space observations. Phys. Rev. Lett. 109 (5), 055001.Google Scholar
Pellat, R., Coroniti, F. V. & Pritchett, P. L. 1991 Does ion tearing exist? Geophys. Res. Lett. 18, 143146.Google Scholar
Petrukovich, A. A., Sergeev, V. A., Zelenyi, L. M., Mukai, T., Yamamoto, T., Kokubun, S., Shiokawa, K., Deehr, C. S., Budnick, E. Y., Büchner, J. et al. 1998 Two spacecraft observations of a reconnection pulse during an auroral breakup. J. Geophys. Res. 103, 4760.Google Scholar
Pitaevskii, L. P 1963 Effect of collisions on perturbation of body rotating in plasma. Sov. J. Expl Theor. Phys. 44, 969979; (in Russian).Google Scholar
Pitaevskii, L. P. & Lifshitz, E. M. 1981 Physical Kinetics, vol. 10. Pergamon.Google Scholar
Porcelli, F., Borgogno, D., Califano, F., Grasso, D., Ottaviani, M. & Pegoraro, F. 2002 Recent advances in collisionless magnetic reconnection. Plasma Phys. Control. Fusion 44, B389.Google Scholar
Quest, K. B., Karimabadi, H. & Brittnacher, M. 1996 Consequences of particle conservation along a flux surface for magnetotail tearing. J. Geophys. Res. 101, 179184.Google Scholar
Rappazzo, A. F., Matthaeus, W. H., Ruffolo, D., Servidio, S. & Velli, M. 2012 Interchange reconnection in a turbulent Corona. Astrophys. J. Lett. 758, L14.CrossRefGoogle Scholar
Schindler, K. 1974 A theory of the substorm mechanism. J. Geophys. Res. 79, 28032810.Google Scholar
Schindler, K. 2006 Physics of Space Plasma Activity. Cambridge University Press.Google Scholar
Servidio, S., Dmitruk, P., Greco, A., Wan, M., Donato, S., Cassak, P. A., Shay, M. A., Carbone, V. & Matthaeus, W. H. 2011 Magnetic reconnection as an element of turbulence. Nonlinear Process. Geophys. 18, 675695.Google Scholar
Sitnov, M. I., Sharma, A. S., Guzdar, P. N. & Yoon, P. H. 2002 Reconnection onset in the tail of Earth’s magnetosphere. J. Geophys. Res. 107, 1256.Google Scholar
Sitnov, M. I. & Swisdak, M. 2011 Onset of collisionless magnetic reconnection in two-dimensional current sheets and formation of dipolarization fronts. J. Geophys. Res. 116, 12216.Google Scholar
Voitenko, Y. M. 1998 Excitation of kinetic Alfvén waves in a flaring loop. Sol. Phys. 182, 411430.Google Scholar
Yoon, P. H. & Lui, A. T. Y. 2006 Quasi-linear theory of anomalous resistivity. J. Geophys. Res. 111, 2203.Google Scholar
Zelenyi, L. & Artemyev, A. 2013 Mechanisms of spontaneous reconnection: from magnetospheric to fusion plasma. Space Sci. Rev. 178, 441457.Google Scholar
Zelenyi, L. M., Artemyev, A. V., Malova, H. V. & Popov, V. Y. 2008 Marginal stability of thin current sheets in the Earth’s magnetotail. J. Atmos. Sol. Terr. Phys. 70, 325333.Google Scholar
Zelenyi, L. M. & Taktakishvili, A. L. 1981 The influence of dissipative processes on the development of the tearing mode in current sheets. Sov. J. Plasma Phys. 7, 10641075.Google Scholar
Figure 0

Figure 1. Schematic view of the X-line region.

Figure 1

Figure 2. Factor $X$ as a function of $L_{x}$ for two values of $B_{z}$ and four values of electron temperature $T$.

Figure 2

Figure 3. Two cases of Cluster spacecraft crossing a reconnection region (see the characteristic reversal of $B_{z}$). Both cases are taken from Artemyev et al. (2015).