Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-25T20:37:54.932Z Has data issue: false hasContentIssue false

Theory of gradient drift instabilities in low-temperature, partially magnetised plasmas

Published online by Cambridge University Press:  15 August 2022

Kentaro Hara*
Affiliation:
Department of Aeronautics and Astronautics, Stanford University, Stanford, CA 94305, USA
Adnan R. Mansour
Affiliation:
Department of Aeronautics and Astronautics, Stanford University, Stanford, CA 94305, USA
Sedina Tsikata
Affiliation:
Centre National de la Recherche Scientifique (CNRS), ICARE UPR 3021, 45071 Orléans, France
*
Email address for correspondence: kenhara@stanford.edu
Rights & Permissions [Opens in a new window]

Abstract

A fluid dispersion theory in partially magnetised plasmas is analysed to examine the conditions under which large-wavelength modes develop in Penning-type configurations, that is, where an electric field is imposed perpendicular to a homogeneous magnetic field. The fluid dispersion relation assuming a slab geometry shows that two types of low-frequency, gradient drift instabilities occur in the direction of the $\boldsymbol {E} \times \boldsymbol {B}$ and diamagnetic drifts. One type of instability, observed when the equilibrium electric field and plasma density gradient are in the same direction, is similar to the classic modified Simon–Hoh instability. A second instability is found for conditions in which (i) the diamagnetic drift is in the direction opposite to the $\boldsymbol {E} \times \boldsymbol {B}$ drift and (ii) the magnitude of the diamagnetic drift is sufficiently larger than the electron thermal speed. The present fluid dispersion theory suggests that the rotating spokes driven by such fluid instabilities propagate in the same direction as the diamagnetic drift, which can be in the same direction as or opposite to the $\boldsymbol {E} \times \boldsymbol {B}$ drift, depending on the plasma conditions. This finding may account for the observation, in some plasma devices, of the rotation of large-scale structures in both the $\boldsymbol {E} \times \boldsymbol {B}$ and $-\boldsymbol {E} \times \boldsymbol {B}$ directions.

Type
Research Article
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), 2022. Published by Cambridge University Press

1. Introduction

Low-temperature magnetised plasmas can be found in a variety of applications and natural phenomena, including magnetron discharges (Keidar & Beilis Reference Keidar and Beilis2006; Anders Reference Anders2012; Ito, Young & Cappelli Reference Ito, Young and Cappelli2015; Hecimovic & von Keudell Reference Hecimovic and von Keudell2018), Penning discharges (Quraishi, Robertson & Walch Reference Quraishi, Robertson and Walch2002; Abolmasov Reference Abolmasov2012), Hall effect thrusters (HETs) and accelerators (Ellison, Raitses & Fisch Reference Ellison, Raitses and Fisch2012; Sekerak et al. Reference Sekerak, Longmier, Gallimore, Brown, Hofer and Polk2015; Romadanov et al. Reference Romadanov, Smolyakov, Raitses, Kaganovich, Tang and Ryzhkov2016; Mazouffre et al. Reference Mazouffre, Grimaud, Tsikata, Matyash and Schneider2019), high-power microwaves (Lau Reference Lau2001; Benford, Swegle & Schamiloglu Reference Benford, Swegle and Schamiloglu2015), dusty (complex) plasmas (Bal & Bose Reference Bal and Bose2010) and interplanetary and interstellar environments (Breneman et al. Reference Breneman, Cattell, Kersten, Paradise, Schreiner, Kellogg, Goetz and Wilson2013). An applied magnetic field traps the charged particles, thus increasing ionisation efficiency and reducing the diffusivity, serving as a critical path to control flows and chemistry in low-temperature plasmas.

Depending on the magnetic field strength and collision frequency, plasmas can be either partially magnetised (i.e. electrons are magnetised but ions are non-magnetised) or fully magnetised (i.e. both ions and electrons are magnetised). In low-temperature plasmas, the dynamics of the neutral gas also play an important role in the transport properties and time-dependent plasma behaviour. The multiscale nature of the plasma flow in such low-temperature magnetised plasmas leads to various oscillation modes from high frequency (of the order of gigahertz) to low frequency (of the order of kilohertz) (Hara Reference Hara2019). Recent studies suggest that the anomalous electron transport across magnetic fields can be due to plasma–wall interaction (Kaganovich et al. Reference Kaganovich, Raitses, Sydorenko and Smolyakov2007; Sydorenko et al. Reference Sydorenko, Smolyakov, Kaganovich and Raitses2008), the plasma waves initiated by kinetic instabilities (Adam, Héron & Laval Reference Adam, Héron and Laval2004; Héron & Adam Reference Héron and Adam2013; Boeuf Reference Boeuf2017; Janhunen et al. Reference Janhunen, Smolyakov, Sydorenko, Jimenez, Kaganovich and Raitses2018) or a combination of these effects (Héron & Adam Reference Héron and Adam2013; Villafana et al. Reference Villafana, Petronio, Denig, Jimenez, Eremin, Garrigues, Taccogna, Alvarez-Laguna, Boeuf and Bourdon2021). A few examples of the kinetic instabilities in the high-frequency range include the electron cyclotron drift instability (Forslund, Morse & Nielson Reference Forslund, Morse and Nielson1971), modified two-stream instability (McBride et al. Reference McBride, Ott, Boris and Orens1972), and ion–ion two-stream instability (Gary Reference Gary1991; Tsikata et al. Reference Tsikata, Cavalier, Héron, Honoré, Lemoine, Grésillon and Coulette2014). On the other hand, low-frequency plasma oscillations include breathing mode oscillations (Boeuf & Garrigues Reference Boeuf and Garrigues1998; Barral & Ahedo Reference Barral and Ahedo2009; Hara et al. Reference Hara, Sekerak, Boyd and Gallimore2014a, Reference Hara, Sekerak, Boyd and Gallimoreb; Dale & Jorns Reference Dale and Jorns2019) and azimuthally rotating spokes (Ellison et al. Reference Ellison, Raitses and Fisch2012; Sekerak et al. Reference Sekerak, Longmier, Gallimore, Brown, Hofer and Polk2015; Kawashima, Hara & Komurasaki Reference Kawashima, Hara and Komurasaki2018). The coupling between different instabilities in various spatial and temporal scales plays an important role in determining the transport coefficients and turbulent phenomena in such low-temperature magnetised plasmas.

In this paper, we derive the dispersion relation of low-frequency, large-wavelength gradient drift instabilities, which may lead to the self-organising patterns in the low-temperature magnetised plasmas. We consider a Penning-type configuration, where a homogeneous axial magnetic field is applied and an electric field and density gradient exist in the radial direction. Although the spokes are often observed in numerical simulations (Boeuf Reference Boeuf2014; Powis et al. Reference Powis, Carlsson, Kaganovich, Raitses and Smolyakov2018; Boeuf & Takahashi Reference Boeuf and Takahashi2020) and in experiments (Raitses, Kaganovich & Smolyakov Reference Raitses, Kaganovich and Smolyakov2015; Marcovati, Ito & Cappelli Reference Marcovati, Ito and Cappelli2020), the mechanism of the spoke formation is still not well understood. One of the most accepted theories attributes the formation of rotating spokes in a Penning-type discharge to the class of Simon–Hoh instabilities (SHIs) (Simon Reference Simon1963; Hoh Reference Hoh1963). This terminology was first used by Sakawa et al. (Reference Sakawa, Joshi, Kaw, Chen and Jain1993), who proposed the modified Simon–Hoh instability (MSHI), which comes from a dispersion relation for partially magnetised plasma, whereas the original work by Simon and Hoh individually focused on fully magnetised plasmas. In work by Simon (Reference Simon1963), the criterion for rotating spoke formation has been proposed to be $\boldsymbol {E}_0\boldsymbol {\cdot } \boldsymbol {\nabla } n_0 >0$, where $\boldsymbol {E}_0$ is the applied electric field and $\boldsymbol {\nabla } n_0$ is the equilibrium plasma density gradient. Recent experimental studies in low-temperature magnetised plasma sources have shown, however, that the rotation direction and speed can depend on various plasma parameters, such as the current (Anders & Yang Reference Anders and Yang2017), indicating that the instability criterion needs to be revisited.

The theory of gradient-drift instabilities in low-temperature magnetised plasmas is introduced in this paper. Section 2 discusses the linearised fluid equations for magnetised plasmas under an applied magnetic field, assuming that the gyroviscosity effects are negligible. Section 3 shows a low-frequency plasma dispersion relation of the partially magnetised plasma considering a slab geometry. The criteria for the large-wavelength, low-frequency, gradient drift instability under a homogeneous magnetic field are discussed. Section 4 illustrates the results of the dispersion relation for partially magnetised plasmas, showing results consistent with the instability criteria derived in § 3. The gradient drift instability theory is applied to various cross-field plasma devices in § 5.

2. Linearised plasma equations for low-temperature magnetised plasmas

Although the rotating spokes are observed in cylindrical (axisymmetric) systems, here we simplify the dispersion relation assuming a slab (Cartesian) geometry. A static, uniform magnetic field in $z$ direction $\boldsymbol {B} = B_0 \hat {\boldsymbol {z}}$ and an equilibrium electric field (applied electric field) are considered: $\boldsymbol {E}_0 = E_0 \hat {\boldsymbol {x}}$. It is assumed that the equilibrium plasma is quasineutral and a plasma density gradient exists locally in the $x$ direction: $E_{0} \neq 0$ and ${\rm d}n_0 /{{\rm d} x} \neq 0$, generating $\boldsymbol {E}\times \boldsymbol {B}$ and diamagnetic drifts in the ${\pm }y$ direction, for the equilibrium condition. In addition, the following assumptions are made: (i) although the temperature gradient may affect the instabilities, we assume that the temperature is uniform and constant, for simplicity; (ii) only ions and electrons are assumed for the species; (iii) the plasma is assumed to be partially magnetised (i.e. magnetised electrons and unmagnetised ions); (iv) collisions are assumed negligible; and (v) the plasma is electrostatic, that is, effects of the induced magnetic field are negligible compared with the applied magnetic field.

2.1. Governing equations

The fluid equations are used for both ions and electrons, i.e. when kinetic effects, such as the Bernstein modes (Bernstein Reference Bernstein1958), can be neglected.

Conservation of mass can be constructed by taking the zeroth moment of the kinetic equation. Here, ionisation and recombination are neglected. Thus, the continuity equation can be written as

(2.1)\begin{equation} \frac{\partial n_s}{\partial t} + \boldsymbol{\nabla}\boldsymbol{\cdot}(n_s\boldsymbol{u}_s) = 0, \end{equation}

where $n_s$ is the number density and $\boldsymbol {u}_s$ is the bulk velocity for species $s$.

The equation for the fluid momentum can be formulated by taking the first moment of the kinetic equation, which can be written using conservative or primitive variables. Assuming that the plasma is collisionless and the distribution function is close to an isotropic Maxwellian distribution function, the conservation of momentum can be written as

(2.2)\begin{equation} \frac{\partial (m_sn_s\boldsymbol{u}_s)}{\partial t} + \boldsymbol{\nabla} \boldsymbol{\cdot}(m_sn_s\boldsymbol{u}_s\boldsymbol{u_s}) ={-}\boldsymbol{\nabla} p_s + q_s n_s(\boldsymbol{E} + \boldsymbol{u}_s\times\boldsymbol{B}), \end{equation}

where $m_s$ is the mass, $p_s$ is the pressure, $q_s$ is the charge, $\boldsymbol {E}$ is the electric field and $\boldsymbol {B}$ is the magnetic field. Using the source-less continuity equation, as shown in (2.1), the momentum equation can also be given, using the primitive variables, by

(2.3)\begin{equation} \frac{\partial \boldsymbol{u}_s}{\partial t} + (\boldsymbol{u}_s\boldsymbol{\cdot} \boldsymbol{\nabla})\boldsymbol{u}_s ={-}\frac{\boldsymbol{\nabla} p_s}{m_s n_s} + \frac{q_s}{m_s} (\boldsymbol{E} + \boldsymbol{u_s}\times \boldsymbol{B}). \end{equation}

Note that the pressure is a scalar term, which is valid when the velocity distribution function (VDF) is close to an isotropic Maxwellian distribution function, that is, the temperatures in three directions are equal. Under this condition, the pressure can be written using the ideal gas law: $p_s = n_s k_B T_s$, where $k_B$ is the Boltzmann constant and $T_s$ is the temperature for species $s$.

In the present model, the ideal gas law is assumed for the electron fluid model. It is to be noted that standard drift models account for gyroviscosity effects, which arise due to the non-Maxwellian distribution and lead to cancellation of the diamagnetic drift in the inertia term in the momentum equation (Ramos Reference Ramos2005; Schnack et al. Reference Schnack, Barnes, Brennan, Hegna, Held, Kim, Kruger, Pankin and Sovinec2006). Although the gyroviscosity effects may play an important role in low-temperature, partially magnetised plasmas (Smolyakov et al. Reference Smolyakov, Chapurin, Frias, Koshkarov, Romadanov, Tang, Umansky, Raitses, Kaganovich and Lakhin2016), the validity of the drift models and the necessity of including gyroviscosity effects in low-temperature magnetised plasmas needs to be investigated. In the state-of-the-art computational models for low-temperature plasmas, simplified fluid models such as the drift-diffusion model are known to represent the physical processes (Kushner Reference Kushner2009; Hara Reference Hara2019). In this paper, a five-moment model that neglects the gyroviscosity effects is used, based on recent numerical simulations that show low-frequency rotating spokes (Mansour & Hara Reference Mansour and Hara2022). The inclusion of the gyroviscosity effects is reserved for future work.

2.2. Linear perturbation analysis for partially magnetised plasmas

We consider the growth of the instabilities in $y$ direction, that is, the direction in which the electrons drift. Under the linear perturbation analysis, a plasma property $Q$ can be described as a sum of the steady-state quantity and a linear perturbation, such that

(2.4)\begin{equation} Q = Q_0 + Q_1 \exp(-{\rm i}\omega t + {\rm i} k_y y), \end{equation}

where $Q_0$ and $Q_1$ are the equilibrium (steady-state) and first-order perturbation terms of a plasma property $Q$, respectively, $\omega$ is the frequency, $t$ is time and $k_y$ is the wave number in the $y$ direction. Here, $\omega = \omega _r + {\rm i}\gamma$, where $\omega _r$ is the real frequency and $\gamma$ is the imaginary part which corresponds to the growth rate.

2.3. Zeroth-order (equilibrium) equations for magnetised electrons

Inserting (2.4) into the governing equations and considering the zeroth-order terms leads to the equilibrium equations. For magnetised electrons, using (2.3) and considering the equilibrium bulk velocity $\boldsymbol {u}_{e0} = (u_{e0x}, u_{e0y}, u_{e0z})^\intercal$, where subscripts $x$, $y$ and $z$ denote the direction, the steady-state momentum equation in $x$ and $y$ directions can be written as

(2.5)\begin{gather} 0 ={-}\frac{k_B T_e}{m_e n_0}\frac{\partial n_0}{\partial x} -\frac{e}{m_e} (E_0 + u_{e0y}B_0) , \end{gather}
(2.6)\begin{gather}0 = \frac{e}{m_e} u_{e0x}B_0 , \end{gather}

where $e$ is the elementary charge, $T_e$ is the electron temperature, $m_e$ is the electron mass and $n_0$ is the equilibrium density, assuming quasineutrality for the equilibrium condition.

Recall that the equilibrium density gradient and electric field are considered to exist locally only in $x$ direction. Although (2.6) results in $u_{e0x}=0$, (2.5) yields the well-known drifts:

(2.7)\begin{equation} u_{e0y} ={-}\frac{E_0}{B_0} - \frac{k_B T_e}{e n_0} \frac{ n_0'}{B_0}, \end{equation}

where $n_0'={\rm d}n_0/{{\rm d} x}$ is the plasma density gradient. The first term in (2.7) is the $\boldsymbol {E} \times \boldsymbol {B}$ drift and the second term is the diamagnetic drift, which can be written as $u_E$ and $u_*$, respectively. The diamagnetic drift does not come from the single-particle trajectory analysis but appears as an equilibrium drift from the fluid theory, whereas the $\boldsymbol {E} \times \boldsymbol {B}$ drift can be derived from single particle trajectories. Nonetheless, the diamagnetic drift is a steady-state bulk velocity that can propagate in the same or opposite direction of the $\boldsymbol {E} \times \boldsymbol {B}$ drift.

As the density gradient is only considered in $x$ direction and using $u_{e0x} = 0$ obtained from (2.6), the steady-state conservation of mass can be written as

(2.8)\begin{equation} \boldsymbol{\nabla}\boldsymbol{\cdot} \boldsymbol{u}_{e0} = 0. \end{equation}

This relation shows that $u_{e0y}$ is constant in $y$ direction, i.e. homogeneous, which is consistent with the configuration considered.

2.4. First-order (linear perturbation) equations for magnetised electrons

Let us consider the perturbation terms of the electron bulk velocity to be $\boldsymbol {u}_{e1} = (u_{e1x}, u_{e1y}, 0)^\intercal$ for magnetised electrons. Using (2.8), the first-order conservation of mass can be derived from (2.1) as

(2.9)\begin{equation} \frac{ \partial n_{e1}}{\partial t} + n_0\boldsymbol{\nabla}\boldsymbol{\cdot} \boldsymbol{u}_{e1} + \boldsymbol{u}_{e0}\boldsymbol{\cdot}\boldsymbol{\nabla} n_{e1} + \boldsymbol{u}_{e1}\boldsymbol{\cdot} \boldsymbol{\nabla} n_0 = 0. \end{equation}

Using the linear perturbation shown in (2.4), the perturbed electron density can be written as

(2.10)\begin{equation} n_{e1} = \frac{n_0k_yu_{e1y} -{\rm i} u_{e1x} n_0'}{\tilde{\omega}}, \end{equation}

where $\tilde {\omega }= \omega - k_yu_{e0y}$. Note that $u_{e0y}$ is given in (2.7).

The first-order momentum equation can be derived from (2.2). Note that pressure term leads to

(2.11)\begin{equation} \frac{k_B T_e \boldsymbol{\nabla}(n_0+n_{e1})}{m_e(n_0+n_{e1})} =\frac{k_B T_e}{m_e}\left[ \frac{\boldsymbol{\nabla} n_0}{n_0} \left( 1 - \frac{n_{e1}}{n_0} \right) + \frac{\boldsymbol{\nabla} n_{e1}}{n_0} \right], \end{equation}

which results in two contributions to the linearised momentum equation. Thus, using (2.11), the first-order momentum equation for magnetised electrons can be given by

(2.12)\begin{equation} -{\rm i}\tilde{\omega} \boldsymbol{u}_{e1} = v_{{\rm th}}^2 k_n \frac{n_{e1}}{n_0} \hat{\boldsymbol{x}} - v_{{\rm th}}^2 \frac{\boldsymbol{\nabla} n_{e1}}{n_0} - \frac{e}{m_e}(\boldsymbol{E}_1 + \boldsymbol{u}_{e1}\times\boldsymbol{B}_0), \end{equation}

where $v_{{\rm th}} = (k_B T_e/m_e)^{1/2}$ is the electron thermal speed and $k_n = n_0' / n_0$ is the inverse of the density gradient length scale, which is also defined similarly by Sakawa et al. (Reference Sakawa, Joshi, Kaw, Chen and Jain1993) and Smolyakov et al. (Reference Smolyakov, Chapurin, Frias, Koshkarov, Romadanov, Tang, Umansky, Raitses, Kaganovich and Lakhin2016). Taking the perturbation terms in $y$ direction, e.g. $\phi _1 \exp (-{\rm i}\omega t +{\rm i} k_y y)$ for the electric field using the electrostatic assumption: $\boldsymbol {E}=-\boldsymbol {\nabla } \phi$, (2.12) can be written for $x$ and $y$ directions as

(2.13)\begin{equation} \begin{bmatrix} -{\rm i}\tilde{\omega} & \omega_{{\rm ce}}\\ -\omega_{{\rm ce}} & -{\rm i}\tilde{\omega} \end{bmatrix} \begin{bmatrix} u_{e1x}\\ u_{e1y} \end{bmatrix} =\begin{bmatrix} v_{{\rm th}}^2 k_n \frac{n_{e1}}{n_0}\\ -{\rm i} v_{{\rm th}}^2 k_y \frac{n_{e1}}{n_0} +{\rm i} \frac{e}{m_e}k_y\phi_1 \end{bmatrix}, \end{equation}

where $\omega _{{\rm ce}} = eB_0/m_e$ is the electron gyrofrequency. Solving for $u_{e1x}$ and $u_{e1y}$ in (2.13) gives

(2.14)\begin{gather} u_{e1x} = \frac{{\rm i}}{\tilde{\omega}^2-\omega_{{\rm ce}}^2} \left[ \tilde{\omega} k_n v_{{\rm th}}^2 \frac{n_{e1}}{n_0} - \omega_{{\rm ce}}k_y\left(v_{{\rm th}}^2 \frac{n_{e1}}{n_0} - \frac{e}{m_e} \phi_1 \right) \right], \end{gather}
(2.15)\begin{gather}u_{e1y} = \frac{1}{\tilde{\omega}^2-\omega_{{\rm ce}}^2} \left[- \omega_{{\rm ce}}k_n v_{{\rm th}}^2 \frac{n_1}{n_0} + \tilde{\omega}k_y\left( v_{{\rm th}}^2 \frac{n_{e1}}{n_0} - \frac{e}{m_e}\phi_1 \right) \right]. \end{gather}

Equations (2.14) and (2.15) are similarly derived in the Rayleigh–Taylor instability analysis (Chen Reference Chen1984). If cold electrons are assumed (i.e. $v_{{\rm th}} = 0$) and one considers the low-frequency approximation (i.e. $\tilde {\omega }^2 \ll \omega _{{\rm ce}}^2$), (2.14) and (2.15) reduce to $u_{e1x} = E_{1y}/B_0$ and $u_{e1y}= {\rm i}\tilde {\omega } u_{e1x}/\omega _{{\rm ce}}$, which are equivalent to the perturbed $\boldsymbol {E} \times \boldsymbol {B}$ and polarisation drifts, respectively. However, it can be seen that the determinant of the matrix in (2.13) becomes negative if $\tilde {\omega }^2 \gg \omega _{{\rm ce}}^2$. In this case, $u_{e1x}$ moves in the opposite direction of the perturbed $\boldsymbol {E} \times \boldsymbol {B}$ drift, which seems to be non-physical (see Appendix A for some discussions). Therefore, assuming $\tilde {\omega }^2 \ll \omega _{{\rm ce}}^2$ and using (2.14) and (2.15), (2.10) can be written as

(2.16)\begin{equation} \frac{n_{e1}}{n_0} = \frac{e\phi_1 }{m_e} \frac{ k_y^2 \tilde{\omega}- k_nk_y \omega_{{\rm ce}}}{ \omega_{{\rm ce}}^2 \tilde{\omega} + \left( k_y^2 + k_n^2\right)v_{{\rm th}}^2 \tilde{\omega} - 2k_nk_y \omega_{{\rm ce}}v_{{\rm th}}^2}. \end{equation}

Note that this equation is similar to (10) in Sakawa et al. (Reference Sakawa, Joshi, Kaw, Chen and Jain1993), except for the coefficient of the last term in the denominator.

2.5. Dispersion function for unmagnetised ions

Let us derive the zeroth-order equations for unmagnetised ions. If restricting the argument to the cross-field direction ($x$ and $y$), that is, neglecting the plasma dynamics in $z$ direction (along the magnetic field), the steady-state continuity equation gives, $n_0 u_{i0x} ={\rm const}.$, where $u_{i0x}$ is the equilibrium ion bulk velocity in $x$ direction. Assuming that the ion bulk velocity is negligible in $y$ direction (i.e. $u_{i0y} = 0$), the linear perturbation of the ion number density, $n_{i1}$, can be derived from the linearised conservation of mass as

(2.17)\begin{equation} n_{i1} = \frac{n_0k_yu_{i1y} -{\rm i} u_{i1x} n_0'}{ \omega - {\rm i} k_n u_{i0x}}, \end{equation}

where ${\boldsymbol {u}}_{i1} = (u_{i1x}, u_{i1y},0)^\intercal$ is the linear perturbation of the ion bulk velocities.

Assuming cold ions ($T_i= 0$, where $T_i$ is the ion temperature) and only considering perturbation in $y$ direction, as shown in (2.4), the linearised momentum equation for unmagnetised ions can be written as

(2.18)\begin{gather} u_{1x} = 0, \end{gather}
(2.19)\begin{gather}u_{1y} = \frac{k_y}{{\omega}} \frac{e}{m_i}\phi_1, \end{gather}

where $m_i$ is the ion mass. If one further assumes that the effects of $k_n u_{i0x}$ to be small, (2.17) can be reduced to

(2.20)\begin{equation} \frac{n_{i1}}{n_0} = \frac{e k_y^2}{m_i \omega^2}\phi_1. \end{equation}

This is consistent with the dispersion function used for unmagnetised ions in Sakawa et al. (Reference Sakawa, Joshi, Kaw, Chen and Jain1993).

3. Gradient drift instability for partially magnetised plasmas

3.1. Fluid dispersion relation

Assuming quasineutrality and perturbations occur in the azimuthal direction, which are similar assumptions employed for the Rayleigh–Taylor instability theory (Chen Reference Chen1984), and using (2.16) and (2.20), the dispersion relation for partially magnetised plasmas can be derived as

(3.1)\begin{equation} \frac{m_e k_y}{m_i \omega^2 } = \frac{ k_y \tilde{\omega}- k_n \omega_{{\rm ce}}}{ [\omega_{{\rm ce}}^2 + \left( k_y^2 + k_n^2\right)v_{{\rm th}}^2] \tilde{\omega} - 2k_nk_y \omega_{{\rm ce}}v_{{\rm th}}^2}, \end{equation}

where $\tilde {\omega } = \omega - k_y u_{e0y}$ is used throughout the derivation. Here, this drift-shifted frequency can be written as $\tilde {\omega } = \omega - \omega _E - \omega _*,$ where $\omega _E = k_y u_E$, $\omega _* = k_y u_*$, $u_E = - E_0 / B_0$ and $u_* = - k_n k_B T_e / (e B_0)$, as can be seen from (2.7). It can be seen that (3.1) yields a third-order equation for $\omega$, from which the damping and linear instability growth can be evaluated.

3.2. Instability criteria in the large-wavelength limit

In the limit of a large-wavelength mode, a simplified form of (3.1) can be obtained by neglecting the $k_y \tilde {\omega }$ term in the numerator of (3.1) compared with $k_n \omega _{{\rm ce}}$, that is, $k_y \tilde {\omega } \ll k_n \omega _{{\rm ce}}$. In addition, $u_* \omega _{{\rm ce}} = - k_n v_{{\rm th}}^2$ gives $k_n^2 v_{{\rm th}}^2 = \tilde {u}_*^2 \omega _{{\rm ce}}^2$, where $\tilde {u}_*=u_*/v_{{\rm th}}$. Therefore, in this limit, (3.1) can be written as

(3.2)\begin{equation} 0 ={-} \frac{k_n m_i }{m_e \omega_{{\rm ce}}} \omega^2 - k_y(1 + \tilde{u}_*^2) \omega + \tilde{u}_*^2 k_y^2 (u_E+u_*) + k_y^2 (u_E-u_*). \end{equation}

For the solution to have an unstable mode, that is, a root with a positive growth rate, the discriminant of (3.2) must be negative. The instability condition can therefore be given by

(3.3)\begin{equation} \frac{m_e }{m_i}(1+\tilde{u}_*^2)^2-4\tilde{u}_* \left[ (1+\tilde{u}_*^2) \tilde{u}_E - (1-\tilde{u}_*^2) \tilde{u}_* \right] < 0, \end{equation}

where $\tilde {u}_E = u_E / v_{{\rm th}}$. By rewriting (3.3) considering $m_e \ll m_i$, the condition for the partially magnetised plasma to have an unstable mode in the limit of large wavelength can be written as

(3.4)\begin{equation} \tilde{u}_* F(\tilde{u}_*, \tilde{u}_E) >0, \end{equation}

where

(3.5)\begin{equation} F(\tilde{u}_*, \tilde{u}_E) = \tilde{u}_*^3 + \tilde{u}_E \tilde{u}_*^2 - \tilde{u}_* + \tilde{u}_E . \end{equation}

As can be seen from (3.4), the two instability criteria can be obtained as (i) $\tilde {u}_*>0$ and $F(\tilde {u}_*, \tilde {u}_E) >0$ and (ii) $\tilde {u}_*<0$ and $F(\tilde {u}_*, \tilde {u}_E) <0$.

It can be seen that (3.3), in the limit of $|\tilde {u}_*| \rightarrow 0$ and considering $m_e \ll m_i$, reduces to $u_E u_* > 0$, which is the instability criterion for the MSHI (Sakawa et al. Reference Sakawa, Joshi, Kaw, Chen and Jain1993). See Appendix B for further discussion about the comparison of the present dispersion relation and MSHI.

Figure 1 shows the shape of (3.5) for two values of $\tilde {u}_E$, that is, $\tilde {u}_E = -0.1$ and $-0.6$, to illustrate the range of unstable and stable roots in the limit of large wavelength and low frequency. Here, negative $u_E$ values are chosen, as the $\boldsymbol {E} \times \boldsymbol {B}$ drift occurs in $-y$ direction when considering $E_0$ in $x$ direction and $B_0$ in $z$ direction. Unlike the MSHI that predicts one region for unstable modes, that is, $\tilde {u}_* \tilde {u}_E > 0$, two unstable regions can be seen in figure 1(a), whereas three unstable regions can be seen in figure 1(b).

Figure 1. Cubic function, $F(\tilde {u}_*, \tilde {u}_E)$, shown in (3.5) for (a) $u_E = -0.6 v_{{\rm th}}$ and (b) $u_E = -0.1 v_{{\rm th}}$, resulting in two and three instability regions, respectively. Instabilities occur according to (3.4) in the regions where the product of the $u_*$ and the function $F(\tilde {u}_*, \tilde {u}_E)$ is positive. The red dashed line denotes $F=0$ and the unstable regions are indicated in blue.

Figure 2 shows the unstable and stable regions of the gradient drift instability in the large-wavelength limit, as a function of the diamagnetic drift and $\boldsymbol {E} \times \boldsymbol {B}$ drift. The unstable roots are obtained according to the instability condition shown in (3.4), and the results are consistent with figure 1, showing the two and three unstable regions depending on $u_*$ and $u_E$. A few observations about the instability criteria can be made as follows.

  1. (i) When $|\tilde {u}_E| > 0.3$, there are two regions that satisfy the instability condition shown in (3.4). One is $u_E u_*>0$, leading to a condition that can be written as $\boldsymbol {E}_0 \boldsymbol {\cdot }\boldsymbol {\nabla } n_0 >0$, which is similar to the MSHI. In this regime, both $\boldsymbol {E} \times \boldsymbol {B}$ and diamagnetic drifts are in the same direction, leading to instability. The other region that results in instability can be found at $u_* > v_{{\rm th}}$ while $u_E < 0$, as shown in figures 1(a) and 2. It is interesting to note that $\boldsymbol {E}_0 \boldsymbol {\cdot }\boldsymbol {\nabla } n_0<0$ in this second region, which is a different instability criterion compared to the MSHI. This indicates that a strong diamagnetic drift (compared with the magnitude of the $\boldsymbol {E} \times \boldsymbol {B}$ drift and the electron thermal speed) can drive an instability at large wavelength whereas the diamagnetic drift is in the opposite direction to the $\boldsymbol {E} \times \boldsymbol {B}$ drift.

  2. (ii) When $0< |\tilde {u}_E| \lessapprox 0.3$, there are three regions that result in linear instability. As can be seen from figure 1(b), $F$ becomes positive in a certain range within $u_*<0$, leading to a stable root because $u_* F < 0$. The appearance of a stable mode in the $\boldsymbol {E}_0 \boldsymbol {\cdot } \boldsymbol {\nabla } n_0 >0$ region can also be seen from figure 2. It can be shown from (3.5) that $F\approx (\tilde {u}_*-\tilde {u}_E)(\tilde {u}_*+ \tilde {u}_E - 1 ) (\tilde {u}_*+ \tilde {u}_E + 1 )$ when assuming $|\tilde {u}_E| \ll 1$. Hence, the three instability regions can be observed approximately at (i) $\tilde {u}_* < -1 - \tilde {u}_E$, (ii) $\tilde {u}_E<\tilde {u}_* < 0$ and (iii) $\tilde {u}_* > 1-\tilde {u}_E$ for $\tilde {u}_E<0$.

  3. (iii) When $\tilde {u}_E=0$, there are two regions where the unstable roots exist. Equation (3.5) reduces to $F = \tilde {u}_* (\tilde {u}_* - 1 ) (\tilde {u}_* + 1 )$ when $\tilde {u}_E=0$. Hence, the instability regions can be observed at (i) $\tilde {u}_* < -1$, and (ii) $\tilde {u}_* > 1$, which can be seen from figure 2. The solutions indicate that a low-frequency, large-wavelength instability can occur with a strong diamagnetic drift in the absence of any electric field.

Figure 2. Unstable and stable regions of the gradient drift instability in the large-wavelength limit, as a function of $u_E$ and $u_*$. The blue region indicates where the unstable modes exist, whereas unstable modes do not exist in the stable region, shown in white.

Figure 3 shows the schematic of the instability criteria for the low-frequency rotating spokes. The first condition $u_E u_*>0$ is similar to the MSHI, in which it is predicted that the $\boldsymbol {E} \times \boldsymbol {B}$ and diamagnetic drifts must be in the same direction. Another instability condition observed from the derivations shown in this section (i.e. (3.4) and (3.5)) illustrates that the partially magnetised plasma can be unstable when the diamagnetic drift is sufficiently larger than the electron thermal speed, even if the electric field and the density gradient exist in the opposite direction.

Figure 3. Instability criteria for low-frequency rotating spokes. (a) The $\boldsymbol {E} \times \boldsymbol {B}$ drift and diamagnetic drift are in the same direction. (b) The $\boldsymbol {E} \times \boldsymbol {B}$ drift and diamagnetic drift are in the opposite direction and the diamagnetic drift must be larger than the electron thermal speed.

3.3. Direction of the wave propagation

When the instability condition is met in (3.3), the real part of the solution in (3.2) can be obtained as

(3.6)\begin{equation} \omega_r = \frac{(1+\tilde{u}_*^2) c_s }{2u_*} k_y c_s, \end{equation}

where $c_s=(k_B T_e /m_i)^{1/2}$ is the ion acoustic speed. The phase velocity of the wave can be obtained from $v_\phi = \omega _r / k_y$. It can therefore be seen from (3.6) that the wave propagates in the direction of the diamagnetic drift, which may lead to the rotating spokes to propagate in the direction of $\boldsymbol {E} \times \boldsymbol {B}$ drift or $-\boldsymbol {E} \times \boldsymbol {B}$ drift, depending on the instability criteria discussed in § 3.2.

3.4. Resonance

It can be seen from (2.14) and (2.15) that the linear perturbation of the electron bulk velocities shows resonance when $\tilde {\omega }^2 \approx \omega _{{\rm ce}}^2$. In other words, $\omega \approx \pm \omega _{{\rm ce}} + k_y u_{e0y}$, where $u_{e0y}$ is the sum of the $\boldsymbol {E} \times \boldsymbol {B}$ drift and diamagnetic drift, given in (2.7). If the real frequency is of the order of ion plasma frequency, $\omega _r \ll \omega _{{\rm ce}}$, as shown later in § 4, a resonance condition for the dispersion relation can be considered to be

(3.7)\begin{equation} k_y ={\mp} \frac{\omega_{{\rm ce}}}{u_{e0y}}. \end{equation}

It can be seen that the smaller (larger) the electron drift, the larger (smaller) $k_y$ at which resonance occurs. Equation (3.7) indicates that the electron drift is in resonance with the electron gyromotion, which is akin to the electrostatic two-stream instability where the electron drift is in resonance with the electron plasma frequency.

Equation (3.7) can also be written as $k_y \lambda _D = \tilde {u}_{e0y}^{-1}{\omega _{{\rm ce}}}/{\omega _{{\rm pe}}}$ or $k_y r_L = \tilde {u}_{e0y}^{-1}$, where $\omega _{{\rm pe}}=(e^2n_0 / m_e \epsilon _0)^{1/2}$ is the electron plasma frequency, $\lambda _D = (\epsilon _0 k_B T_e / e^2 n_0)^{1/2}$ is the Debye length, $\epsilon _0$ is the vacuum permittivity, $\tilde {u}_{e0y}=u_{e0y}/v_{{\rm th}}$ and $r_L = v_{{\rm th}}/\omega _{{\rm ce}}$ is the electron Larmor radius. Although the kinetic effects, such as the electron Bernstein mode leading to electron cyclotron drift instability (Cavalier et al. Reference Cavalier, Lemoine, Bonhomme, Tsikata, Honoré and Grésillon2013; Hara & Tsikata Reference Hara and Tsikata2020), may play an important role when $k_y r_L=O(1)$, the results in the present fluid theory are applicable for the smaller $k_y$ range, e.g. $k_y r_L<1$. Investigation of the coupling of the present fluid theory and the kinetic dispersion relation (Chang & Callen Reference Chang and Callen1992) is reserved for future work.

4. Results

In this section, the results from the linear perturbation theory are discussed. Although the analytic discussions made in § 3.2 are applicable only in the large-wavelength, low-frequency limit, the dispersion relation introduced in (3.1) is valid for a wide range of $k_y$ under the assumption of the fluid approach and before the resonance condition as shown in (3.7). It is discussed in this section that the growth rate exhibits a broadband profile and the real frequency is of the order of the ion plasma frequency, which suggests that the wave propagation speed is of the order of the ion acoustic speed.

The following conditions are considered for two cases: $B_0= 200$ G, $n_0=10^{15}\ {\rm m}^{-3}$, the electron temperature is $4$ eV and $m_i=40$ amu (assuming singly charged argon ions). These plasma parameters are representative of low-temperature cross-field plasma sources. Here, the two cases considered are (I) $E_0 = 10\ {\rm kV}\ {\rm m}^{-1}$ and (II) $E_0 = 4\ {\rm kV}\ {\rm m}^{-1}$. Cases I and II result in $u_E \approx -0.6 v_{{\rm th}}$ and $u_E \approx -0.24 v_{{\rm th}}$, respectively, which correspond to the two and three instability region cases as shown in figure 1. The solution to (3.1) is evaluated for several representative values of $u_*$.

4.1. Case I: two instability regions, $u_E \approx -0.6 v_{{\rm th}}$

Figure 4 shows the real and imaginary parts of the solution of (3.1) for different $u_*$ values for $u_E \approx -0.6 v_{{\rm th}}$. This is consistent with the two instability region case, which occurs when $|u_E| > 0.3$, as shown in figure 1(a).

Figure 4. Case I: two instability regions at large wavelength, for different values of $u_*$ with $u_E$ fixed at $u_E = -0.6 v_{{\rm th}}$. This value of $u_E = -0.6 v_{{\rm th}}$ corresponds to the case illustrated in figure 1(a). The real frequency, $\omega _r$, and growth rate, $\gamma$, are shown in blue dashed and black solid lines, respectively, for each value of $u_*$.

In the limit of a large positive diamagnetic drift (e.g. $u_* = 1.67 v_{{\rm th}}$ as shown in figure 4a), a low-frequency mode is indeed observed in the limit of large wavelength. The real frequency, $\omega _r$, is positive, indicating that wave propagates in the direction of the diamagnetic drift as opposed to the $\boldsymbol {E} \times \boldsymbol {B}$ drift. This observation is consistent with the theoretical observation in § 3.2.

Figures 4(b) and 4(c) show that the dispersion relation yields roots with a positive growth rate at a finite $k_y\neq 0$ but yields stable solutions in the large-wavelength limit, as the magnitude of the diamagnetic drift decreases. Note that for these cases, $u_E u_*<0$, that is, the diamagnetic drift is in the direction opposite to the $\boldsymbol {E} \times \boldsymbol {B}$ drift. These results are consistent with the prediction from the analytic theory obtained in § 3.2. It is to be noted that the sign of the real frequency flips, which can be seen from figures 4(b) and 4(c), indicating that the resonance type phenomenon is driven by the diamagnetic drift in figure 4(b). However, the wave propagates in the direction of the $\boldsymbol {E} \times \boldsymbol {B}$ drift in figure 4(c) despite the fact that the magnitude of diamagnetic drift is larger than the $\boldsymbol {E} \times \boldsymbol {B}$ drift. As the diamagnetic drift approaches zero, the resonance condition shifts towards a larger $k_y \lambda _D$, as shown in (3.7), and the partially magnetised plasma is stable for a wide range of $k_y$.

When the diamagnetic drift is in the same direction as the $\boldsymbol {E} \times \boldsymbol {B}$ drift (i.e. $u_E u_*>0$), the results in figure 4(df) illustrate that the large-wavelength mode is indeed unstable. This is consistent with § 3.2, in which it is observed that $u_E u_*>0$ yields a root with a positive growth rate at the large-wavelength, low-frequency limit, if $|u_E|> 0.3 v_{{\rm th}}$. Another observation that can be made is that the real frequency is negative, illustrating that the wave propagates in the direction of both $\boldsymbol {E} \times \boldsymbol {B}$ drift and diamagnetic drift. The real frequency where the growth rate is maximum is approximately $(5\text {-}6)\omega _{{\rm pi}}$, suggesting that the wave propagation speed is of the order of the ion acoustic speed. The cutoff where $\gamma$ becomes zero at a finite $k_y$ in figure 4(df) occurs before the resonance condition discussed in § 3.4.

4.2. Case II: three instability regions, $u_E \approx -0.24 v_{{\rm th}}$

Figure 5 shows the real and imaginary parts of the solution of (3.1) for several representative $u_*$ values for $u_E \approx -0.24 v_{{\rm th}}$. As discussed in § 3.2 and shown in figure 1(b), three instability regions can be seen in the large-wavelength limit, when $|u_E| \lessapprox 0.3 v_{{\rm th}}$.

Figure 5. Case II: three instability regions at large wavelength, for different values of $u_*$ with $u_E$ fixed at $u_E = -0.24 v_{{\rm th}}$. The real frequency, $\omega _r$, and growth rate, $\gamma$, are shown in blue dashed and black solid lines, respectively, for each value of $u_*$. In figure 5(e), the growth rate is zero at $k_y \lambda _D \ll 1$, which corresponds to the region where $u_*<0$ and $F>0$, shown in figure 1(b).

Similar to Case I, there is a low-frequency mode at large wavelength when the diamagnetic drift is opposite to the $\boldsymbol {E} \times \boldsymbol {B}$ drift and the magnitude of the diamagnetic drift is sufficiently larger than the electron thermal speed, as shown in figure 5(a). Under this situation, the wave propagation occurs in the direction of the diamagnetic drift. As the magnitude of the diamagnetic drift is decreased, the partially magnetised plasma becomes stable in the large-wavelength limit but a resonance type mode appears, as shown in figures 5(b) and 5(c), similar to Case I. In addition, when the diamagnetic drift is in the same direction as the $\boldsymbol {E} \times \boldsymbol {B}$ drift, the partially magnetised plasma is unstable in the large-wavelength limit, which can be seen in figures 5(d) and 5f). However, it is interesting to note that the growth rate becomes zero at $k_y \rightarrow 0$ for the cases with intermediate $u_*$ values, as shown in figure 5(e). This is consistent with the theoretical prediction in § 3.2 and figure 1(b): when $|u_E| \lessapprox 0.3 v_{{\rm th}}$, the partially magnetised plasma becomes stable at $k_y \rightarrow 0$ for a range of $u_*$ in which $u_* F<0$, as shown in figure 1(b).

In summary, the partially magnetised plasma dispersion relation using the fluid approach shows that the linear growth rate is positive in the large-wavelength limit not only (i) when $\boldsymbol {E}_0 \boldsymbol {\cdot } \boldsymbol {\nabla } n_0 >0$ but also (ii) when $\boldsymbol {E}_0 \boldsymbol {\cdot } \boldsymbol {\nabla } n_0 <0$ and the diamagnetic drift is sufficiently larger than the electron thermal speed. Although the former is similar to the so-called MSHI, the latter suggests that a strong diamagnetic drift may excite large-wavelength, low-frequency plasma oscillations that propagate in the direction of the diamagnetic drift.

5. Application of the gradient drift instability to cross-field plasma sources

Now that we have established the theory of the gradient drift instability, in this section, the theory is applied to various cross-field plasma configurations. Although the theory discussed in the present paper assumes a slab geometry and the real systems are cylindrical, the slab approximation may provide first-order estimates for the local instabilities in cross-field plasma sources. Here, four configurations are considered: (a) Penning discharge, (b) cylindrical magnetron (e.g. for high-power microwave generation), (c) planar magnetron discharge (e.g. for plasma-assisted deposition) and (d) Hall effect thruster (HET), as shown in figure 6.

Figure 6. Application of the gradient drift instability theory to various cross-field plasma devices. Note that the geometries between top and bottom figures are rotated to keep the magnetic field direction consistent, that is, into the page. Blue arrows, that is, $u_{0r}$ in (a,b) or $u_{0z}$ in (c,d), indicate the cross-field electron flow, whereas the red arrows indicate the electron drift in the azimuthal direction.

5.1. Penning discharge

The Penning discharge, as shown in figure 6(a), operates using an outer cylinder as an anode with a cathode placed along the centreline, in addition to an applied axial magnetic field (Hoh Reference Hoh1963; Simon Reference Simon1963). The plasma density is typically largest near the cathode, generating a radially inward plasma density gradient, $\boldsymbol {\nabla }_r n_0<0$. At the same time, the applied electric field is also radially inward, $E_r<0$. Penning discharge typically naturally satisfies the condition $\boldsymbol {E}_0 \boldsymbol {\cdot }\boldsymbol {\nabla } n_0 >0$, that is, the $\boldsymbol {E} \times \boldsymbol {B}$ and diamagnetic drifts are in the same direction. Therefore, as shown in figure 3(a), the Penning-type discharge satisfies (3.4), generating a gradient drift instability at low frequency.

5.2. Cylindrical magnetron

The cylindrical magnetron, shown in figure 6(b), is used to study critical ionisation velocity (CIV) phenomena (Brenning et al. Reference Brenning, Lundin, Minea, Costin and Vitelaru2013). In addition, such a cross-field configuration is used in high-power microwave sources. For its use as a microwave generator, ion formation is to be avoided because the working principle is that the electrons from the cathode are trapped (i.e. insulated) by the applied or induced magnetic field to generate high-power microwaves (Benford et al. Reference Benford, Swegle and Schamiloglu2015). However, as the devices increase in power and become more compact, plasma generation is unavoidable as the current density in the system increases (Hadas et al. Reference Hadas, Sayapin, Krasik, Bernshtam and Schnitzer2008). Although low-frequency plasma oscillations are not well-studied due to the short pulse operation, it can be seen from figure 6(b) that the partially magnetised plasma is unstable in such configurations, similar to the Penning discharge.

5.3. Planar magnetron discharge

Figure 6(c) shows a schematic of the planar magnetron, used for plasma-assisted sputtering and deposition (Waits Reference Waits1978). An axial electric field and a radial magnetic field are applied close to the cathode surface. The crossed electric and magnetic fields generate electron drifts in the azimuthal direction.

Recent experimental observations in high-power impulse magnetron sputtering (HiPIMS) show that the rotating spoke direction may be a function of the current (Anders & Yang Reference Anders and Yang2017) and can be reversed during a certain operation (Hecimovic et al. Reference Hecimovic, Maszl, von der Gathen, Böke and von Keudell2016). Similar reversal of the rotating spoke propagation is found in a micro-magnetron discharge (Ito et al. Reference Ito, Young and Cappelli2015; Marcovati et al. Reference Marcovati, Ito and Cappelli2020). The global axial plasma profile, such as the location and amplitude of the density and potential gradients, can affect the characteristics of the gradient drift instability, leading to rotating spokes.

In addition, magnetron sputtering devices have a large magnetic field inhomogeneity in the region where the rotating spokes occur, leading to the possibility of gradient drift instabilities due to magnetic field gradients (Esipchuk & Tilinin Reference Esipchuk and Tilinin1976; Tilinin Reference Tilinin1977; Frias et al. Reference Frias, Smolyakov, Kaganovich and Raitses2013) playing an important role in the formation of rotating spokes.

5.4. HET

The HET operates using a radial applied magnetic field and axial electric field, as shown in figure 6(d). In the plume of a HET, the $\boldsymbol {E} \times \boldsymbol {B}$ drift is much larger than the diamagnetic drift and the diamagnetic drift is in the opposite direction to the $\boldsymbol {E} \times \boldsymbol {B}$ drift, which is shown as region 2 in figure 6(d). Consequently, using the instability criteria discussed in (3.4), the partially magnetised plasma in the HET plume is stable against the Penning-type instability. The simulation results by Kawashima et al. (Reference Kawashima, Hara and Komurasaki2018) were compared with a class of gradient drift instabilities in the presence of a magnetic field gradient (Esipchuk & Tilinin Reference Esipchuk and Tilinin1976; Tilinin Reference Tilinin1977; Frias et al. Reference Frias, Smolyakov, Kaganovich and Raitses2013) and showed that the gradient drift instability due to the magnetic field gradient grows in the plume region, which leads to the excitation of rotating spokes.

However, if one considers the plasma inside the channel, shown as region 1 in figure 6(d), the cross-field electron transport is primarily driven by the pressure gradient near the anode. It can therefore be considered that the magnitude of the diamagnetic drift is larger than that of the $\boldsymbol {E} \times \boldsymbol {B}$ drift, while the two drifts may be in the opposite direction. Thus, this condition may lead to the gradient drift instability presented in this paper to be unstable. The spoke rotation will be in the direction of the diamagnetic drift driven by the plasma density gradient inside the channel, which turns out to be in the same direction as the $\boldsymbol {E} \times \boldsymbol {B}$ drift due to the applied electric field (set up by the anode and cathode) in the system.

In summary, there are two scenarios in which azimuthally rotating spokes can occur in HETs and in the planar magnetron discharge. One is due to the instabilities that are caused by the magnetic field gradient. The other possibility is the gradient drift instability, discussed in the present paper, which is caused by the diamagnetic drift in the absence of magnetic field gradients.

6. Conclusion

The present paper has reviewed the fluid dispersion relation of the partially magnetised plasmas. The dispersion relation derived in this paper shows that the partially magnetised plasmas are unstable (i) when $\boldsymbol {E}_0\boldsymbol {\cdot }\boldsymbol {\nabla } n_0 >0$ or (ii) in the presence of a large diamagnetic drift while $\boldsymbol {E}_0\boldsymbol {\cdot }\boldsymbol {\nabla } n_0 <0$. The former is consistent with the so-called MSHI where the $\boldsymbol {E} \times \boldsymbol {B}$ drift and the diamagnetic drift are in the same direction. The latter is an instability that occurs when the diamagnetic drift is sufficiently larger than the electron thermal speed while occurring in the direction against the $\boldsymbol {E} \times \boldsymbol {B}$ drift. One similarity between the two instability conditions is that the plasma wave in the large-wavelength limit propagates in the direction of the diamagnetic drift. This indicates that the low-frequency, large-wavelength, partially magnetised plasma oscillations can occur in either $\boldsymbol {E} \times \boldsymbol {B}$ or $-\boldsymbol {E} \times \boldsymbol {B}$ direction, depending on the plasma conditions.

Acknowledgements

This material is based on work supported by the US Department of Energy, Office of Science, Office of Fusion Energy Sciences, under Award No. DE-SC0020623, by the Air Force Office of Scientific Research under Awards No. FA9550-18-1-0090 and No. FA9550-21-1-0433, and by the Office of Naval Research under Award No. N00014-21-1-2698.

Editor Edward Thomas, Jr. thanks the referees for their advice in evaluating this article.

Declaration of interests

The authors report no conflict of interest.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Authors contributions

K.H. derived the theory and held discussions with A.R.M. and S.T. Results were prepared by K.H. and all authors contributed to analysing data, reaching conclusions and in writing the paper.

Appendix A. Effect of $\tilde {\omega }^2$ in the electron bulk velocity dispersion

In (2.16), it is assumed that $\tilde {\omega }^2 \ll \omega _{{\rm ce}}^2$. In the limit of large $k_y$, the fluid dispersion relation, shown in (3.1), can be written as

(A1)\begin{equation} 0 = ( \omega^2- k_y^2 c_{s}^2 )\tilde{\omega} . \end{equation}

The solution to the dispersion relation yields three roots, that is, $\omega = \pm k_y c_s, k_y u_{e0y}$. Thus, there are no growth rates in the limit of $k_y \rightarrow \infty$ in figures 4 and 5.

If one retains the $\tilde {\omega }^2$ term in (2.14) and (2.15), the electron density perturbation term can be written as,

(A2)\begin{equation} \frac{n_{e1}}{n_0} = \frac{e\phi_1 }{m_e} \frac{ k_y^2 \tilde{\omega}- k_nk_y \omega_{{\rm ce}}}{ -\tilde{\omega}^3 + \omega_{{\rm ce}}^2 \tilde{\omega} + \left( k_y^2 + k_n^2\right)v_{{\rm th}}^2 \tilde{\omega} - 2k_nk_y \omega_{{\rm ce}}v_{{\rm th}}^2}. \end{equation}

Combining (A2) and (2.20), a third-order equation can be constructed for $\omega$. For this case, the dispersion relation in the limit of $k_y \rightarrow \infty$ can be given by

(A3)\begin{equation} 0 = \left( \omega^2- k_y^2 c_{s}^2 + \frac{m_e }{m_i} \tilde{\omega}^2 \right)\tilde{\omega} . \end{equation}

In (A3), one solution is always real, that is, $\omega =k_y u_{e0y}$. However, the other two solutions must be evaluated separately. Using $m_e \ll m_i$, (A3) can be rewritten as

(A4)\begin{equation} 0 = \omega^2 - 2\frac{m_e }{m_i} k_y u_{e0y} \omega - k_y^2 \left(c_{s}^2 + \frac{m_e }{m_i}u_{e0y}^2 \right) . \end{equation}

The discriminant of (A4) can be obtained as

(A5)\begin{equation} D= \left( \frac{m_e }{m_i} k_y u_{e0y} \right)^2 - k_y^2 \left(c_{s}^2 + \frac{m_e }{m_i}u_{e0y}^2 \right), \end{equation}

which can be seen to be negative (i.e. $D<0$), because $m_e \ll m_i$. Thus, these two solutions are not real and become imaginary at $k_y \rightarrow \infty$.

Although the inclusion of the $\tilde {\omega }^3$ term affects the results in the larger $k_y$ region, the large-wavelength solutions ($k_y \lambda _D < 1$) shown in figures 4 and 5 are not affected by the inclusion of the $\tilde {\omega }^3$ term. Therefore, in the present paper, the derivations and results neglecting the $\tilde {\omega }^3$ term in (A2) are shown.

Appendix B. Comparison with MSHI theory

The condition for gradient drift instability shown in (3.4) and (3.5) is different from the condition that is originally proposed by Simon (Simon Reference Simon1963), that is, $\boldsymbol {E}_0 \boldsymbol {\cdot }\boldsymbol {\nabla } n_0 >0$.

To derive a dispersion relation similar to the MSHI (Sakawa et al. Reference Sakawa, Joshi, Kaw, Chen and Jain1993), $|\tilde {u}_*|\ll 1$ and $k_y\rightarrow 0$ can be assumed so that (3.1) reduces to

(B1)\begin{equation} \frac{k_y^2 c_s^2}{ \omega^2 } = \frac{ \omega_*}{ \omega - \omega_E + \omega_*}, \end{equation}

where $c_s = (k_B T_e / m_i)^{1/2}$ is the ion acoustic speed. Assuming $m_e \ll m_i$, the instability criterion from (B1) can be derived as $u_*(u_E - u_*)>0,$ which leads to $|u_E|>|u_*|$ and $u_E u_* >0$. This means that, assuming $|\tilde {u}_*|\ll 1$, a positive growth rate can be obtained only when the magnitude of diamagnetic drift is larger than that of $\boldsymbol {E}\times \boldsymbol {B}$ drift and $\boldsymbol {E}_0 \boldsymbol {\cdot }\boldsymbol {\nabla } n_0 >0$ are both satisfied.

It should be noted that the dispersion relation proposed by Sakawa et al. (Reference Sakawa, Joshi, Kaw, Chen and Jain1993) omitted a factor of two in the last term of the denominator of the right-hand side in (3.1). This results in a dispersion relation that omits the $\omega _*$ term in the denominator in (B1), leading to the instability condition to be unconditionally $u_* u_E >0$ and therefore $\boldsymbol {E}_0 \boldsymbol {\cdot } \boldsymbol {\nabla } n_0 >0$. It is discussed in § 3.2 that the MSHI can be recovered in the limit of $|\tilde {u}_*| \rightarrow 0$, which can be seen from (B1) as well.

References

REFERENCES

Abolmasov, S.N. 2012 Physics and engineering of crossed-field discharge devices. Plasma Sources Sci. Technol. 21 (3), 035006.10.1088/0963-0252/21/3/035006CrossRefGoogle Scholar
Adam, J.C., Héron, A. & Laval, G. 2004 Study of stationary plasma thrusters using two-dimensional fully kinetic simulations. Phys. Plasmas 11 (1), 295305.10.1063/1.1632904CrossRefGoogle Scholar
Anders, A. 2012 Self-organization and self-limitation in high power impulse magnetron sputtering. Appl. Phys. Lett. 100 (22), 224104.10.1063/1.4724205CrossRefGoogle Scholar
Anders, A. & Yang, Y. 2017 Direct observation of spoke evolution in magnetron sputtering. Appl. Phys. Lett. 111 (6), 064103.10.1063/1.4994192CrossRefGoogle Scholar
Bal, S. & Bose, M. 2010 Modified Simon–Hoh instability in a magnetized inhomogeneous dusty plasma. J. Phys.: Conf. Ser. 208, 012081.Google Scholar
Barral, S. & Ahedo, E. 2009 Low-frequency model of breathing oscillations in Hall discharges. Phys. Rev. E 79, 046401.10.1103/PhysRevE.79.046401CrossRefGoogle ScholarPubMed
Benford, J., Swegle, J.A. & Schamiloglu, E. 2015 High Power Microwaves, 3rd edn. CRC.10.1201/b19681CrossRefGoogle Scholar
Bernstein, I.B. 1958 Waves in a plasma in a magnetic field. Phys. Rev. 109, 1021.10.1103/PhysRev.109.10CrossRefGoogle Scholar
Boeuf, J.-P. 2014 Rotating structures in low temperature magnetized plasmas—insight from particle simulations. Front. Phys. 2, 74.10.3389/fphy.2014.00074CrossRefGoogle Scholar
Boeuf, J.-P. 2017 Tutorial: physics and modeling of Hall thrusters. J. Appl. Phys. 121 (1), 011101.10.1063/1.4972269CrossRefGoogle Scholar
Boeuf, J.-P. & Garrigues, L. 1998 Low frequency oscillations in a stationary plasma thruster. J. Appl. Phys. 84 (7), 35413554.10.1063/1.368529CrossRefGoogle Scholar
Boeuf, J.-P. & Takahashi, M. 2020 Rotating spokes, ionization instability, and electron vortices in partially magnetized $E\times B$ plasmas. Phys. Rev. Lett. 124, 185005.10.1103/PhysRevLett.124.185005CrossRefGoogle Scholar
Breneman, A.W., Cattell, C.A., Kersten, K., Paradise, A., Schreiner, S., Kellogg, P.J., Goetz, K. & Wilson, L.B. III 2013 Stereo and wind observations of intense cyclotron harmonic waves at the Earth's bow shock and inside the magnetosheath. J. Geophys. Res.: Space 118 (12), 76547664.10.1002/2013JA019372CrossRefGoogle Scholar
Brenning, N., Lundin, D., Minea, T., Costin, C. & Vitelaru, C. 2013 Spokes and charged particle transport in HiPIMS magnetrons. J. Phys. D: Appl. Phys. 46 (8), 084005.10.1088/0022-3727/46/8/084005CrossRefGoogle Scholar
Cavalier, J., Lemoine, N., Bonhomme, G., Tsikata, S., Honoré, C. & Grésillon, D. 2013 Hall thruster plasma fluctuations identified as the $E\times B$ electron drift instability: modeling and fitting on experimental data. Phys. Plasmas 20 (8), 082107.10.1063/1.4817743CrossRefGoogle Scholar
Chang, Z. & Callen, J.D. 1992 Unified fluid/kinetic description of plasma microinstabilities. Part I: basic equations in a sheared slab geometry. Phys. Fluids B: Plasma 4 (5), 11671181.10.1063/1.860125CrossRefGoogle Scholar
Chen, F.F. 1984 Introduction to Plasma Physics and Controlled Fusion. Springer.10.1007/978-1-4757-5595-4CrossRefGoogle Scholar
Dale, E.T. & Jorns, B.A. 2019 Non-invasive time-resolved measurements of anomalous collision frequency in a Hall thruster. Phys. Plasmas 26 (1), 013516.10.1063/1.5077008CrossRefGoogle Scholar
Ellison, C.L., Raitses, Y. & Fisch, N.J. 2012 Cross-field electron transport induced by a rotating spoke in a cylindrical Hall thruster. Phys. Plasmas 19 (1), 013503.10.1063/1.3671920CrossRefGoogle Scholar
Esipchuk, Y.B. & Tilinin, G.N. 1976 Drift instability in a Hall-current plasma accelerator. Sov. Phys. Tech. Phys. 21, 417.Google Scholar
Forslund, D.W., Morse, R.L. & Nielson, C.W. 1971 Nonlinear electron-cyclotron drift instability and turbulence. Phys. Rev. Lett. 27 (21), 1424.10.1103/PhysRevLett.27.1424CrossRefGoogle Scholar
Frias, W., Smolyakov, A.I., Kaganovich, I.D. & Raitses, Y. 2013 Long wavelength gradient drift instability in Hall plasma devices. II. Applications. Phys. Plasmas 20 (5), 052108.10.1063/1.4804281CrossRefGoogle Scholar
Gary, S.P. 1991 Electromagnetic ion/ion instabilities and their consequences in space plasmas: a review. Space Sci. Rev. 56 (3), 373415.10.1007/BF00196632CrossRefGoogle Scholar
Hadas, Y., Sayapin, A., Krasik, Ya.E., Bernshtam, V. & Schnitzer, I. 2008 Plasma dynamics during relativistic S-band magnetron operation. J. Appl. Phys. 104 (6), 064125.10.1063/1.2986520CrossRefGoogle Scholar
Hara, K 2019 An overview of discharge plasma modeling for Hall effect thrusters. Plasma Sources Sci. Technol. 28, 044001.10.1088/1361-6595/ab0f70CrossRefGoogle Scholar
Hara, K., Sekerak, M.J., Boyd, I.D. & Gallimore, A.D. 2014 a Mode transition of a Hall thruster discharge plasma. J. Appl. Phys. 115 (20), 203304.10.1063/1.4879896CrossRefGoogle Scholar
Hara, K., Sekerak, M.J., Boyd, I.D. & Gallimore, A.D. 2014 b Perturbation analysis of ionization oscillations in Hall effect thrusters. Phys. Plasmas 21 (12), 122103.10.1063/1.4903843CrossRefGoogle Scholar
Hara, K. & Tsikata, S. 2020 Cross-field electron diffusion due to the coupling of drift-driven microinstabilities. Phys. Rev. E 102, 023202.10.1103/PhysRevE.102.023202CrossRefGoogle Scholar
Hecimovic, A., Maszl, C., von der Gathen, V.S., Böke, M. & von Keudell, A. 2016 Spoke rotation reversal in magnetron discharges of aluminium, chromium and titanium. Plasma Sources Sci. Technol. 25 (3), 035001.10.1088/0963-0252/25/3/035001CrossRefGoogle Scholar
Hecimovic, A. & von Keudell, A. 2018 Spokes in high power impulse magnetron sputtering plasmas. J. Phys. D: Appl. Phys. 51 (45), 453001.10.1088/1361-6463/aadaa1CrossRefGoogle Scholar
Héron, A. & Adam, J.C. 2013 Anomalous conductivity in Hall thrusters: effects of the non-linear coupling of the electron-cyclotron drift instability with secondary electron emission of the walls. Phys. Plasmas 20 (8), 082313.10.1063/1.4818796CrossRefGoogle Scholar
Hoh, F.C. 1963 Instability of Penning-type discharges. Phys. Fluids 6 (8), 11841191.10.1063/1.1706878CrossRefGoogle Scholar
Ito, T., Young, C.V. & Cappelli, M.A. 2015 Self-organization in planar magnetron microdischarge plasmas. Appl. Phys. Lett. 106 (25), 254104.10.1063/1.4922898CrossRefGoogle Scholar
Janhunen, A., Smolyakov, A., Sydorenko, D., Jimenez, M., Kaganovich, I. & Raitses, Y. 2018 Evolution of the electron cyclotron drift instability in two-dimensions. Phys. Plasmas 25 (8), 082308.10.1063/1.5033896CrossRefGoogle Scholar
Kaganovich, I.D., Raitses, Y., Sydorenko, D. & Smolyakov, A. 2007 Kinetic effects in a Hall thruster discharge. Phys. Plasmas 14 (5), 057104.10.1063/1.2709865CrossRefGoogle Scholar
Kawashima, R., Hara, K. & Komurasaki, K. 2018 Numerical analysis of azimuthal rotating spokes in a crossed-field discharge plasma. Plasma Sources Sci. Technol. 27 (3), 035010.10.1088/1361-6595/aab39cCrossRefGoogle Scholar
Keidar, M. & Beilis, I.I. 2006 Electron transport phenomena in plasma devices with $E\times B$ drift. IEEE Trans. Plasma Sci. 34 (3), 804814.10.1109/TPS.2006.874852CrossRefGoogle Scholar
Kushner, M.J. 2009 Hybrid modelling of low temperature plasmas for fundamental investigations and equipment design. J. Phys. D: Appl. Phys. 42 (19), 194013.10.1088/0022-3727/42/19/194013CrossRefGoogle Scholar
Lau, Y.Y. 2001 Simple theory for the two-dimensional Child–Langmuir law. Phys. Rev. Lett. 87, 278301.10.1103/PhysRevLett.87.278301CrossRefGoogle ScholarPubMed
Mansour, A.R. & Hara, K. 2022 Full fluid moment modeling of rotating spokes in Penning-type configuration. Plasma Sources Sci. Technol. 31, 055012.10.1088/1361-6595/ac6a73CrossRefGoogle Scholar
Marcovati, A., Ito, T. & Cappelli, M.A. 2020 The dynamics of coherent modes of gradient drift instabilities in a small magnetron discharge plasma. J. Appl. Phys. 127 (22), 223301.10.1063/5.0006320CrossRefGoogle Scholar
Mazouffre, S., Grimaud, L., Tsikata, S., Matyash, K. & Schneider, R. 2019 Rotating spoke instabilities in a wall-less Hall thruster: experiments. Plasma Sources Sci. Technol. 28 (5), 054002.10.1088/1361-6595/ab07fcCrossRefGoogle Scholar
McBride, J.B., Ott, E., Boris, J.P. & Orens, J.H. 1972 Theory and simulation of turbulent heating by the modified two-stream instability. Phys. Fluids 15 (12), 23672383.10.1063/1.1693881CrossRefGoogle Scholar
Powis, A.T., Carlsson, J.A., Kaganovich, I.D., Raitses, Y. & Smolyakov, A. 2018 Scaling of spoke rotation frequency within a Penning discharge. Phys. Plasmas 25 (7), 072110.10.1063/1.5038733CrossRefGoogle Scholar
Quraishi, Q., Robertson, S. & Walch, R. 2002 Electron diffusion in the annular Penning trap. Phys. Plasmas 9 (8), 32643271.10.1063/1.1495873CrossRefGoogle Scholar
Raitses, Y., Kaganovich, I.D. & Smolyakov, A. 2015 Effects of the gas pressure on low frequency oscillations in $E\times B$ discharges. In Proceedings of Joint Conference of 30th International Symposium on Space Technology and Science 34th International Electric Propulsion Conference and 6th Nano-satellite Symposium, pp. IEPC–2015–307.Google Scholar
Ramos, J.J. 2005 General expression of the gyroviscous force. Phys. Plasmas 12 (11), 112301.10.1063/1.2114747CrossRefGoogle Scholar
Romadanov, I., Smolyakov, A., Raitses, Y., Kaganovich, I., Tang, T. & Ryzhkov, S. 2016 Structure of nonlocal gradient-drift instabilities in Hall $E\times B$ discharges. Phys. Plasmas 23 (12), 122111.10.1063/1.4971816CrossRefGoogle Scholar
Sakawa, Y., Joshi, C., Kaw, P.K., Chen, F.F. & Jain, V.K. 1993 Excitation of the modified Simon–Hoh instability in an electron beam produced plasma. Phys. Fluids B: Plasma 5 (6), 16811694.10.1063/1.860803CrossRefGoogle Scholar
Schnack, D.D., Barnes, D.C., Brennan, D.P., Hegna, C.C., Held, E., Kim, C.C., Kruger, S.E., Pankin, A.Y. & Sovinec, C.R. 2006 Computational modeling of fully ionized magnetized plasmas using the fluid approximation. Phys. Plasmas 13 (5), 058103.10.1063/1.2183738CrossRefGoogle Scholar
Sekerak, M.J., Longmier, B.W., Gallimore, A.D., Brown, D.L., Hofer, R.R. & Polk, J.E. 2015 Azimuthal spoke propagation in Hall effect thrusters. IEEE Trans. Plasma Sci. 43 (1), 7285.10.1109/TPS.2014.2355223CrossRefGoogle Scholar
Simon, A. 1963 Instability of a partially ionized plasma in crossed electric and magnetic fields. Phys. Fluids 6 (3), 382388.10.1063/1.1706743CrossRefGoogle Scholar
Smolyakov, A.I., Chapurin, O., Frias, W., Koshkarov, O., Romadanov, I., Tang, T., Umansky, M., Raitses, Y., Kaganovich, I.D. & Lakhin, V.P. 2016 Fluid theory and simulations of instabilities, turbulent transport and coherent structures in partially-magnetized plasmas of $E\times B$ discharges. Plasma Phys. Control. Fusion 59 (1), 014041.10.1088/0741-3335/59/1/014041CrossRefGoogle Scholar
Sydorenko, D., Smolyakov, A., Kaganovich, I. & Raitses, Y. 2008 Plasma-sheath instability in Hall thrusters due to periodic modulation of the energy of secondary electrons in cyclotron motion. Phys. Plasmas 15 (5), 053506.10.1063/1.2918333CrossRefGoogle Scholar
Tilinin, G.N. 1977 High-frequency plasma waves in a Hall accelerator with an extended acceleration zone. Zh. Tekh. Fiz. 47, 16841691.Google Scholar
Tsikata, S., Cavalier, J., Héron, A., Honoré, C., Lemoine, N., Grésillon, D. & Coulette, D. 2014 An axially propagating two-stream instability in the Hall thruster plasma. Phys. Plasmas 21 (7), 072116.10.1063/1.4890025CrossRefGoogle Scholar
Villafana, W., Petronio, F., Denig, A.C., Jimenez, M.J., Eremin, D., Garrigues, L., Taccogna, F., Alvarez-Laguna, A., Boeuf, J.P., Bourdon, A., et al. 2021 2D radial-azimuthal particle-in-cell benchmark for $E\times B$ discharges. Plasma Sources Sci. Technol. 30 (7), 075002.10.1088/1361-6595/ac0a4aCrossRefGoogle Scholar
Waits, R.K. 1978 Planar magnetron sputtering. J. Vac. Sci. Technol. 15 (2), 179187.10.1116/1.569451CrossRefGoogle Scholar
Figure 0

Figure 1. Cubic function, $F(\tilde {u}_*, \tilde {u}_E)$, shown in (3.5) for (a) $u_E = -0.6 v_{{\rm th}}$ and (b) $u_E = -0.1 v_{{\rm th}}$, resulting in two and three instability regions, respectively. Instabilities occur according to (3.4) in the regions where the product of the $u_*$ and the function $F(\tilde {u}_*, \tilde {u}_E)$ is positive. The red dashed line denotes $F=0$ and the unstable regions are indicated in blue.

Figure 1

Figure 2. Unstable and stable regions of the gradient drift instability in the large-wavelength limit, as a function of $u_E$ and $u_*$. The blue region indicates where the unstable modes exist, whereas unstable modes do not exist in the stable region, shown in white.

Figure 2

Figure 3. Instability criteria for low-frequency rotating spokes. (a) The $\boldsymbol {E} \times \boldsymbol {B}$ drift and diamagnetic drift are in the same direction. (b) The $\boldsymbol {E} \times \boldsymbol {B}$ drift and diamagnetic drift are in the opposite direction and the diamagnetic drift must be larger than the electron thermal speed.

Figure 3

Figure 4. Case I: two instability regions at large wavelength, for different values of $u_*$ with $u_E$ fixed at $u_E = -0.6 v_{{\rm th}}$. This value of $u_E = -0.6 v_{{\rm th}}$ corresponds to the case illustrated in figure 1(a). The real frequency, $\omega _r$, and growth rate, $\gamma$, are shown in blue dashed and black solid lines, respectively, for each value of $u_*$.

Figure 4

Figure 5. Case II: three instability regions at large wavelength, for different values of $u_*$ with $u_E$ fixed at $u_E = -0.24 v_{{\rm th}}$. The real frequency, $\omega _r$, and growth rate, $\gamma$, are shown in blue dashed and black solid lines, respectively, for each value of $u_*$. In figure 5(e), the growth rate is zero at $k_y \lambda _D \ll 1$, which corresponds to the region where $u_*<0$ and $F>0$, shown in figure 1(b).

Figure 5

Figure 6. Application of the gradient drift instability theory to various cross-field plasma devices. Note that the geometries between top and bottom figures are rotated to keep the magnetic field direction consistent, that is, into the page. Blue arrows, that is, $u_{0r}$ in (a,b) or $u_{0z}$ in (c,d), indicate the cross-field electron flow, whereas the red arrows indicate the electron drift in the azimuthal direction.