Electron cyclotron drift instability (ECDI) and ion–ion two stream instability (IITSI) are both kinetic instabilities that can be present in low-temperature, partially magnetized plasmas. The coupling of instabilities in a three-dimensional configuration leads to the existence of more than one unstable roots to the kinetic dispersion relation. In this paper, a generalized method has been developed for numerically evaluating solutions to the three-dimensional dispersion relation for coupled ECDI and IITSI, assuming cold singly and doubly charged ions and a Maxwellian velocity distribution function for the electrons. The present study demonstrates the coupling between ECDI and IITSI that affects the most unstable mode as a function of the wavenumbers in three dimensions and various plasma properties, including the applied electric field, magnetic field, electron temperature, ion velocities, and plasma density. One of the most notable results is that, while the most unstable mode with the largest growth rate is in the direction of the E×B drift in the two-dimensional cases, the most unstable mode for the three-dimensional configuration occurs in the oblique direction between the applied electric field and the E×B drift. This agrees with experimental observations in cross field plasma sources using coherent Thomson scattering.

Low-temperature, partially magnetized plasmas are used for a wide range of applications, including material processing, space propulsion, and beam sources. In these plasmas, electrons are magnetized, whereas ions are unmagnetized. While crossed electric and magnetic fields are used to confine electrons, the electron drifts, such as the E×B drift (where E is the applied electric field, and B is the applied magnetic field), cause plasma instabilities, leading to plasma turbulence and anomalous electron transport across magnetic fields.

The main kinetic instabilities that occur in partially magnetized plasmas include the electron cyclotron drift instability (ECDI)1,2 and modified two-stream instability (MTSI)3,4 driven by the electron E×B drift. In laboratory cross field discharge sources, coherent Thomson scattering experiments have shown the existence of plasma waves driven by such kinetic instabilities in the direction of E×B drift.5,6 A number of kinetic simulations have recently revisited these instabilities and their effect on cross field electron transport.7–11 In addition to the plasma waves in the E×B direction, Tsikata et al.12 have found experimental evidence of plasma waves that propagate in the direction of the applied electric field (E) resulting from an ion–ion two-stream instability (IITSI) driven by the relative drift velocity between singly and doubly charged ions in the cross field plasma discharge. Hara and Tsikata13 further observed using a particle-in-cell (PIC) simulation that the coupling of ECDI and IITSI in a two-dimensional (2D) configuration results in microturbulence, which enhances electron diffusion across the magnetic fields. Furthermore, the cross field electron transport is found to be affected by the magnitude and wavelength of the plasma waves.14,15

ECDI occurs due to coupling of electron Bernstein waves and unmagnetized ions. The Bernstein mode in three-dimensional (3D), i.e., when spatial fluctuations are accounted for in the direction parallel and perpendicular to the magnetic field, results in the use of the plasma dispersion function assuming that the electrons follow a Maxwellian velocity distribution function (VDF), while the 1D and 2D assumptions reduce the plasma dispersion function to resonance-type modes that can be described using polynomials.16 Cavalier et al.17 developed a numerical method to obtain the solution of the 3D ECDI dispersion relation for singly charged ions. For such cases, there are two solutions that are complex conjugates, allowing for an iterative method to obtain the most unstable root. However, when multiply charged ion streams are considered, IITSI occurs in addition to ECDI, and the resulting modes are no longer a superposition of the two instabilities. As shown later, when ECDI and IITSI coexist, more than one unstable root exists, making the iterative method infeasible. Thus, a generalized numerical solver to calculate an arbitrary number of unstable roots of a kinetic dispersion relation is needed.

In this paper, we investigate the three-dimensional coupling of ECDI and IITSI in a partially magnetized plasma through the development of a numerical technique for finding complex roots of the coupled dispersion relation. A theoretical examination of ECDI and IITSI and the root-finding algorithm are presented in Sec. II. Section III discusses the dispersion relation for singly charged ions and magnetized electrons, validating the root-finding algorithm with published results. In Sec. IV, the dispersion relation for singly and doubly charged ions and magnetized electrons is presented. Here, the coupled ECDI and IITSI yield more than one unstable root, which the generalized numerical algorithm can obtain.

Plasma instabilities can be discussed in the context of a linear perturbation to the equilibrium distribution function of the form exp[i(ωtk·r)] with ω=ωr+iγ, where ω is the wave frequency, k is the wave vector, t is the time, r is the position vector, ωr is the real frequency, and γ is the growth rate. In general, the electrostatic dispersion relation can be obtained from Poisson's equation as

(1)

where χe is the electron susceptibility, and χi is the ion susceptibility. The susceptibilities can be constructed from the Vlasov equation, assuming a collisionless plasma.

In this study, the plasma is assumed to be partially magnetized, i.e., the electrons are magnetized, and the ions are non-magnetized. It is further assumed that (i) the ions are cold, and (ii) the electrons follow a Maxwellian VDF for the equilibrium condition, from which the linear perturbation evolves. These assumptions follow the physics of cross field discharges, e.g., Hall effect thrusters (HETs). The ions are typically generated via electron-impact ionization from neutral gas, making the plasma non-equilibrium (TiTe, where Ti and Te are the ion and electron temperatures) in the region where E×B drift is largest.18,19 Additionally, while a few numerical investigations have studied the effects of non-Maxwellian electron distributions,20,21 a Maxwellian VDF is often considered for electrons in various computational models22,23 and linear instability studies.17,24–26 The effects of finite ion temperature (e.g., ion Landau damping27) and non-Maxwellian VDFs on the kinetic instabilities will be reserved for future work.

The assumptions used in this study are consistent with previous ECDI studies.12,13,17 In this section, the 3D (and 2D) ECDI theory is first reviewed to discuss the fixed-point iteration method by Cavalier et al.17,28 Then, the dispersion relation due to the coupling of ECDI and IITSI, i.e., when another ion stream with different velocity exists, is introduced. Finally, the root finding algorithm and results are shown.

Taking into account the electron dynamics in directions parallel and perpendicular to the magnetic field while assuming that the electrons are magnetized and the equilibrium distribution function is a Maxwellian VDF, the electron susceptibility can be written as

(2)

where k=|k|=(k2+k2)1/2,k=(kx2+ky2)1/2 is the wavenumber perpendicular to the magnetic field, k=kz is the wave number parallel to the magnetic field, λD=(ϵ0kBTe/e2n0)1/2 is the Debye length, Ud is the drift velocity for the electrons (later we consider the E×B drift), Ui+ is the singly charged ion bulk velocity, vth,s=(kBTs/ms)1/2 is the thermal velocity of species s=e,i, and ωce=eB0/me is the electron cyclotron frequency. Here, me and mi are the electron and ion masses, ϵ0 is the permittivity of free space, kB is the Boltzmann constant, e is the elementary charge, n0 is the quasineutral plasma density, b=k2rL2,rL=vth,e/ωce is the electron Larmor radius at the thermal velocity, Z is the plasma dispersion function given by Z(ξ)=π1/2exp(t2)(tξ)1dt, and In is the modified Bessel function of the first kind and order n. It is to be noted that the assumption that the electrons follow a Maxwellian VDF for the equilibrium condition results in the well-known electron Bernstein mode,16 as shown in Eq. (2). For non-Maxwellian VDFs, the electron susceptibility needs to be reevaluated.29 

Assuming cold ions (ω/kvth,i), the ion susceptibility term can be rewritten as

(3)

where the ion plasma frequency is given by ωpi=(e2n0/ϵ0mi)1/2.

1. 3D ECDI dispersion

Using Eqs. (1)–(3), the ECDI dispersion relation of cold, singly charged ions, and magnetized electrons in a three-dimensional (3D) configuration17 can be written as

(4)

As shown later in Sec. II A 3, this 3D ECDI dispersion relation has only two roots which are complex conjugates. Thus, a fixed-point iteration scheme works to solve Eq. (4).

Another particular instability that arises from the 3D ECDI dispersion relation is called the modified two-stream instability (MTSI), which exists for finite, non-zero kλD. The dispersion relation for MTSI is derived by taking the fluid limit,3 i.e., krL1, of Eq. (4) in which only the n = 0 term is retained in the sum of the electron susceptibility

(5)

The motion of electrons both parallel and perpendicular to magnetic field lines is taken into account. It is to be noted that this dispersion relation is called the MTSI as it resembles the two-stream instability, in which two species counter stream. The final term in Eq. (5) can be considered that electrons have an effective mass of (k2/k2)me, which can be large for small kλD. Unstable modes exist that are largely perpendicular to the applied magnetic field with k/k(me/mi)1/2 such that the effective electron mass is on the order of the ion mass.3 Because of this fact, these waves cause electron heating parallel to the magnetic field that is on the order of ion heating perpendicular to the magnetic field.9,30

2. 2D ECDI dispersion

In the limit of kλD0, the dispersion relation can be considered to be two-dimensional (in the k direction). The argument of the plasma dispersion function, ξ, approaches infinity, allowing the use of the asymptotic expansion for Z(ξ). Thus, Eq. (4) reduces to

(6)

It can be seen from Eq. (6) that resonances exist for all higher harmonics n=1,2, if (ωk·Ud)2=(nωce)2. In this limit of wave propagation perpendicular to the magnetic field, Ducrocq et al.31 discusses the existence of instability for angular frequencies in the drift frame close to cyclotron resonances, i.e., k·Udnωce for positive integer n. These electron Bernstein waves32 dominate the behavior of ECDI in the 2D (k) limit. Additionally, it can be seen that Eq. (6), i.e., 2D dispersion relation, can be written in a form using polynomials, whereas the 3D dispersion relation requires the solution of the plasma dispersion function as shown in Eq. (4).

In 3D, i.e., for kλD0, the plasma wave along magnetic field lines plays an important role in how the instabilities grow. Smaller wavelength oscillations that propagate parallel to the magnetic field results in a finite kvth,e, and the resonant structure existing for perpendicular propagation broadens, eventually becoming a broadband ion acoustic-like spectrum.13 Resonance broadening occurs as the summation of Bessel functions become smaller perturbations to the ion acoustic-like terms.9 However, it is important to emphasize the distinction between the ion acoustic-like instability in three-dimensional ECDI and the ion acoustic instability in unmagnetized plasmas. The kinetic dispersion relation for ion acoustic waves in an unmagnetized, Maxwellian plasma for vth,iωr/kvth,e is given below:

(7)

Equation (7) can be obtained33 from Eq. (4) when b, i.e., B0.

3. Solution of the 3D ECDI dispersion relation

To calculate the roots of the 3D ECDI dispersion relation shown in Eq. (4), Cavalier et al. developed a numerical algorithm in Ref. 17. This algorithm relies on the structure of the ECDI dielectric function in order to rewrite the dispersion relation in the following form:

(8)

where all terms with a tilde are normalized: k is normalized by the Debye length λD, ω is normalized by the ion plasma frequency ωpi, velocities (Ui+ and Ud) are normalized by the ion acoustic speed cs=kBTe/mi, and M̃=mi/me is the ratio of the ion mass to electron mass.17 The function g(Ω,X,Y) in Eq. (8) is referred to as the Gordeev function34 

(9)

Writing the 3D ECDI dispersion relation in the form of Eq. (8) enables the use of a fixed-point iteration scheme, i.e., the unstable root of Eq. (8) is obtained iteratively until convergence is achieved. To initialize the iteration, ω0=0 is used to evaluate the Gordeev function.17 This scheme leads to two solutions that have different values for ωr (off by k̃·Ũi+) and complex conjugate values for γ, i.e., the magnitude of the imaginary part of the two roots is always identical. This allows one to find the most unstable root iteratively.

The discussion in Sec. II A assumes that all ions are cold, singly charged, and unmagnetized. However, if there is some fraction of cold ions that are doubly charged, i.e., the doubly charged ion temperature is also much smaller than the electron temperature, given by α=2ni2+/ne where ni2+ is the number density of doubly charged ions that carry two positive charges, and ne is the electron number density; then, the contribution of the doubly charged ions to the dispersion relation must be accounted for. With two ion species, the quasineutrality condition is n0=ne=ni++2ni2+. For the sake of consistency, the ion plasma frequency, ωpi, is still defined using n0, the electron number density. Hence, in the presence of a mixture of singly and doubly charged ions, the 3D kinetic dispersion relation can be written as

(10)

where χe is the electron contribution to the dispersion relation, identical to Eq. (2), and Ui2+ is the velocity of the doubly charged ion stream. Equation (10) can be non-dimensionalized with the same quantities as Eq. (8) in the following way:

(11)

where ω̃d is the non-dimensionalized form of ωd=ωk·Ui+=ωd,r+iγ, the angular frequency shifted into the frame moving at the singly charged ion bulk velocity, and ΔŨi is the non-dimensionalized form of ΔUi=Ui2+Ui+, i.e., the difference in the singly and doubly charged bulk velocities. It can be seen that Eq. (11) reduces to Eq. (8) when α = 0, i.e., when doubly charged ions are absent.

While the dispersion relation in the case of singly charged ions and magnetized electrons in Eq. (8) has two solutions that are complex conjugates (i.e., resulting in one unstable root), there are four possible solutions to Eq. (11) leading to two unstable roots. Therefore, the fixed-point iteration algorithm by Cavalier et al.17 struggles to find the most unstable root in the presence of two (or more) unstable roots. As a result, a new approach must be taken for equations of the form given in Eq. (10).

To numerically evaluate kinetic dispersion relations in the form of Eq. (10), the plasma dispersion function must be calculated for complex inputs. First, the method by which the plasma dispersion function is numerically evaluated is verified. The plasma dispersion function can be reformulated as Z(ξ)=iπexp(ξ2)erfc(iξ)=iπw(ξ), where erfc is the complementary error function, and w(ξ) is the Faddeeva function. The Faddeeva function is numerically evaluated using two different methods, depending on the modulus of the argument ξ. For sufficiently large |ξ|, a continued fraction expansion, such as the one reported by Gautschi,35 is used. For smaller |ξ|, the method by Zaghloul and Ali36 is employed, which evaluates the Faddeeva function along a carefully selected integration contour between 0 and z=x+iy. To verify the accuracy of the computation of Z(ξ), a benchmark was performed against the work by Baalrud37 for both real values of ξ[0,10] and complex values of ξ=x+iy for x[0,3.5] and y[3,3].

For the coupled ECDI and IITSI cases, the solution of Eq. (10) can result in two or more roots that have γ>0, corresponding to linear instability growth. Therefore, to study the coupling of ECDI and IITSI, it is necessary to use an algorithm that is capable of enumerating all roots of the dielectric function for given values of λD, Ud,vth,e, ωce, Ui+,Ui2+, α, and k. This is achieved in the following manner: for a given set of values for the aforementioned parameters, the dispersion relation given by Eq. (10) is evaluated for a range of values for ωr and γ. With this information, it is possible to analyze the ϵr(ω)=0 and ϵi(ω)=0 contours for fixed k, where ϵr and ϵi are the real and imaginary components of ϵ(ω,k). Roots of the dielectric function exist for values of ωr and γ which simultaneously lie on the ϵr=0 and ϵi=0 contours.

In order to locate the intersection(s) of the two contours, a standard root-finding algorithm38 is considered. First, for a given condition, ϵr and ϵi are calculated on discrete ωr and γ. Next, the line segments of ϵr=0 and ϵi=0 are obtained. The line segments are iterated over pairwise to check for points of intersection. To eliminate the unnecessary overhead of solving for the point(s) of intersection of line segments which cannot possibly cross, the pairs of endpoints of each line segment are compared to determine if an intersection is possible. If so, a matrix inversion is performed to determine the location of the intersection along the line segments. The runtime of the algorithm is affected by how finely the contours are discretized and the structure of the contours themselves, which is a function of the electric and magnetic fields, the plasma density, electron temperature, ion bulk velocities, k, and α. Acceleration of the root finding algorithm is reserved for future investigation.

For the remainder of the paper, we consider an applied electric field E=E0x̂ and an applied magnetic field B=B0ẑ. Hence, kz=k. This results in an E×B drift that occurs in the −y direction.

Figure 1 illustrates the contours of ϵi=0 (red) and ϵr=0 (black) for three different conditions for a given k: k̃x = 0.5, k̃y = 0.5, and k̃z = k̃ = 0.09. Here, ΔŨi=1.57. ECDI occurs by considering Ũd=117, and IITSI is excited assuming α=0.5. The red and black contours contain 700 and 1000 line segments, respectively. This method of solving dispersion relations is roughly 1000 times more computationally expensive than the fixed-point algorithm of Cavalier et al.17 due to the need of root finding using discretized frequencies (both real frequency and growth rate).

FIG. 1.

Complex contour of ϵ shown in Eq. (10) for different instabilities: ϵr=0 (black) and ϵi=0 (red), for k̃x=0.5,k̃y=0.5,k̃z=k̃=0.09: (a) ECDI-only: α = 0 and Ũd=117, (b) IITSI-only: α=0.5,Ũd=0, and ΔŨi=1.57, and (c) ECDI-IITSI coupled: α=0.5,Ũd=117, and ΔŨi=1.57. The blue circles represent roots of ϵ(ω,k)=0, and the gray dashed lines show where γ and ωd,r=ωrkxUi+ are 0.

FIG. 1.

Complex contour of ϵ shown in Eq. (10) for different instabilities: ϵr=0 (black) and ϵi=0 (red), for k̃x=0.5,k̃y=0.5,k̃z=k̃=0.09: (a) ECDI-only: α = 0 and Ũd=117, (b) IITSI-only: α=0.5,Ũd=0, and ΔŨi=1.57, and (c) ECDI-IITSI coupled: α=0.5,Ũd=117, and ΔŨi=1.57. The blue circles represent roots of ϵ(ω,k)=0, and the gray dashed lines show where γ and ωd,r=ωrkxUi+ are 0.

Close modal

Figure 1(a), i.e., the ECDI-only case, contains four closed curves, two of which comprise the ϵi=0 contour and two of which comprise the ϵr=0 contour. The two solutions of Eq. (10) with α = 0, or, equivalently Eq. (4), denoted by the blue circle in Fig. 1(a), can be seen to be the intersections of ϵi=0 and ϵr=0 contours. These solutions are complex conjugates, for which the fixed-point iteration works. In addition, it can be seen that the real frequencies are symmetric about the ωd,r=ωrkxUi+=0, i.e., ωr=kxUi+ line. Because instability growth exists for γ>0, the solution to the ECDI dispersion relation is the root with ωd,r>0, leading to a formation of plasma waves in the direction of k=(kx,ky,kz).

However, it is to be noted that the curves encompassing the ϵi=0 contour never intersect, and the same is true for the two curves within the ϵr=0 contour. Hence, the curves do not intersect at ωd,r=0 and γ = 0, meaning that this point is not a solution to Eq. (4). Yet, when discretizing the (ωd,r,γ) space, the two contours seem to intersect and these intersections tend to be detected as the solutions to the dispersion relation with the line segment method. As a result, because these numerical intersections exist for very small |γ|/ωpi values, an assumption is made in the root-finding algorithm that only roots with |γ|/ωpi>0.005 are considered as a solution to Eq. (4). This simplification to the root-finding algorithm works well for much of the (kx, ky) space, except in between low k resonances in which the instability growth rates are considerably small. This cutoff should not affect finding the roots of larger growth rate modes which are more physically relevant.

Figure 1(b) shows the IITSI-only case, where Ũd is set to zero and α is set to 0.5. Both the real and imaginary components of ϵ are modified from the ECDI-only case. The addition of the doubly charged ion contribution leads to another lobe for ϵr=0. The pole (note: not a solution) of ϵr=0 and ϵi=0 at ωd,r=γ=0 corresponds to the singly charged ion contribution, while that at ωd,r=k̃xΔŨi=0.785 corresponds to the doubly charged ion contribution. Similar to the ECDI-only case in Fig. 1(a), the complex roots exist only when ϵr=0 and ϵi=0 contours intersect, as shown using the blue circles in Fig. 1(b). Two roots correspond to γ = 0, while the other two roots are complex conjugates. The visualization of the roots is reminiscent of the traditional two-stream instability,16 which exhibits two real roots and two complex conjugate roots if an instability condition is met. For various combinations of parameters, i.e., various wavenumbers and singly and doubly charged bulk velocities, the fastest propagating real root may also become slightly damped [γ/ωpiO(103)]. The complex conjugate roots of the IITSI-only case in Fig. 1(b) have the same value of ωr while the growth rates are same in magnitude but have opposite signs, i.e., symmetric across the γ = 0 line. The ωr is determined mainly by the IITSI resonance, and ωr>0 means that the waves initiated by the IITSI will be in the direction of the k. The real part of the IITSI roots, therefore, lies between the singly and doubly charged ion velocities.13 

Finally, Fig. 1(c) shows the results of a coupled ECDI and IITSI, which have features similar to the ECDI-only [Fig. 1(a)] and IITSI-only [Fig. 1(b)] cases. It can be seen in Fig. 1(c) that ϵr=0 contour is similar to the IITSI-only case while the two ϵi=0 contours (similar to the ECDI-only case) stem from both the singly and doubly charged ion branches. Such a profile yields four roots for the coupled ECDI-IITSI solution. In addition, it can be seen that the ϵ contours are no longer symmetric due to the coupling of ECDI and IITSI. Most importantly, there are two roots with γ>0, and these roots are not complex conjugates. Hence, one needs to quantitatively evaluate both these roots to obtain the most unstable root. Depending on the shape of the ϵ contours, it can be seen from Fig. 1(c) that one unstable root is due to the ϵi=0 contour stemming from the singly charged ion branch and the other is due to the ϵi=0 contour stemming from the doubly charge ion branch. This illustrates the need of a generalized complex root-finding method for arbitrarily complex dispersion relations. This process provides a method for numerically solving dispersion relations by mapping the problem into one which conventional root-finding algorithms can be leveraged without making an assumption about the structure of the (ωr,γ)-solution space. The variation of the dispersion relation as a function of the wavenumber will be shown and discussed later (cf. Fig. 4).

First, the 3D ECDI dispersion relation [Eq. (8)] is solved with only singly charged ions for small kxλD to verify the root-finding algorithm against the fixed-point iteration method developed by Cavalier et al.17 Twenty-one terms (n[10,10]) are kept in the summation given in Eq. (9) such that the maximum truncation error is on the order of 106. The relevant simulation parameters are given in Table I. The angular frequency and growth rate for kzλD=0.01, 0.03, and 0.09 are compared.

TABLE I.

Parameters chosen for the benchmarking with the 3D ECDI dispersion solver by Cavalier et al.17 

Simulation parameters
E0 (V/m) 7520 
B0 (T) 0.015 
Te (eV) 25 
ne (m−32×1017 
Ui+ (m/s) 16×103 
mi (amu) 131 
Simulation parameters
E0 (V/m) 7520 
B0 (T) 0.015 
Te (eV) 25 
ne (m−32×1017 
Ui+ (m/s) 16×103 
mi (amu) 131 

Figure 2 demonstrates excellent agreement between the fixed-point iteration scheme of Cavalier et al.17 [see Eq. (8)] and the complex root-finding algorithm as discussed in Sec. II D. In the singly charged ions only case, the discrete resonances are evident for small kzλD, approaching the 2D limit, as shown in Eq. (6). These resonances occur for ky near the resonance condition kyUdnωce. From the values shown in Table I, ω̃ce=51.2 and Ũd=117.2, corresponding to resonance conditions of kyλD0.437n for n. The second, third, and fourth modes visible in Fig. 2 are consistent with this metric, lying within 0.05ωpi of the predicted E×B harmonic. The lowest frequency mode, which is particularly visible for kzλD=0.01, corresponds to MTSI, as discussed in Eq. (5). Using the MTSI dispersion relation, the maximum growth rate3 is found to be γmax=ωLH/2, where ωLH is the lower-hybrid frequency defined by ωLH=ωpi(1+ωpe2/ωce2)1/2. Using the parameters mentioned above, the normalized lower-hybrid frequency is ω̃LH=0.104, resulting in a maximum growth rate for MTSI of γ̃max=0.052. The solution obtained from the root-finding algorithm is in good agreement with the theoretical prediction of the MTSI growth rate. Furthermore, Gary and Sanderson39 predicts that in the limit of Udωr/k, the wave number of maximum growth of ECDI is kyλD=1/2. When the applied electric field is increased to 104 V/m, the wave number of maximum growth from the 3D dispersion relation is found to be approximately kyλD=0.68, which is in good agreement with Gary's theory. Finally, the resonant structure of ECDI observed for small kzλD transitions into a broadband, ion acoustic-like instability as kzλD increases. This is evident through the broadband signature of the growth rate, as shown in Fig. 2(b), over a wide range of ky values for kzλD0.03. These metrics, in addition to the benchmarking of the other modules in the 3D dispersion solver, demonstrate that the three-dimensional ECDI roots are being properly evaluated.

FIG. 2.

ECDI dispersion relation for singly charged ions only (α = 0 and kxλD=0.01), using the fixed-point iteration scheme (solid lines) and complex root-finding scheme (stars): (a) real frequency ωr and (b) growth rate γ.

FIG. 2.

ECDI dispersion relation for singly charged ions only (α = 0 and kxλD=0.01), using the fixed-point iteration scheme (solid lines) and complex root-finding scheme (stars): (a) real frequency ωr and (b) growth rate γ.

Close modal

The plots of the drift-shifted real frequency ωd,r=ωrkxUi+ and the growth rate γ as functions of kx and ky are illustrated in Fig. 3 using a 180 × 180 grid of kx and ky values. The real frequency plots are masked such that only (kx, ky) values that have a nonzero growth rate are shown. Note that Fig. 2 is identical to the results of Figs. 3(a)–3(c) for kxλD=0.01. The characteristics of the three-dimensional ECDI growth rates largely reflect the same trend discussed in Fig. 2. The notable difference between the three cases is the decrease in the resonance strength as kx increases. As such, the E×B dynamics in the y-direction become less pronounced, and the instability growth rate becomes more broadband as kzλD increases. Similarly, the unstable regions of IITSI, as shown in Fig. 3(d), agree with the work of Hara and Tsikata13 in that unstable regions exist for kxλD<1, the growth rate increases with ky for kxλD[0.4,0.9], and the shifted real frequency monotonically increases with kx. The three-dimensional ECDI and IITSI growth rates for kzλD=0.01 are both roughly 50% off the two-dimensional (kzλD=0) values evaluated by Hara and Tsikata.13 In contrast, the real frequency is similar between the 2D and 3D cases for ECDI and IITSI, further benchmarking the root-finding algorithm for finite kz.

FIG. 3.

ECDI dispersion relation for singly charged ions only (α = 0) as a function of kx and ky: (1) drift-shifted real frequency and (2) the growth rate as a function of kxλD and kyλD for (a) kzλD=0.01, (b) kzλD=0.03, and (c) kzλD=0.09. The IITSI drift-shifted real frequency and the growth rate are also shown for (d) α=0.2 and kzλD=0.01. The solid lines are shown to visualize the contour lines for small values. Maximum values of the colorbar for ECDI: γ/ωpi = 0.23 and ωd,r/ωpi=0.97; and IITSI: γ/ωpi = 0.34 and ωd,r/ωpi=1.30.

FIG. 3.

ECDI dispersion relation for singly charged ions only (α = 0) as a function of kx and ky: (1) drift-shifted real frequency and (2) the growth rate as a function of kxλD and kyλD for (a) kzλD=0.01, (b) kzλD=0.03, and (c) kzλD=0.09. The IITSI drift-shifted real frequency and the growth rate are also shown for (d) α=0.2 and kzλD=0.01. The solid lines are shown to visualize the contour lines for small values. Maximum values of the colorbar for ECDI: γ/ωpi = 0.23 and ωd,r/ωpi=0.97; and IITSI: γ/ωpi = 0.34 and ωd,r/ωpi=1.30.

Close modal

Using the root-finding algorithm, described in Sec. II D and verified in Sec. III, we now investigate the coupling between the ECDI and IITSI due to the mixture of singly and doubly charged ion streams. The results obtained in this section evaluate Eq. (11) with α0 and Ui2+=22.71×103 m/s in addition to the parameters shown in Table I. Using these conditions, ΔŨi=O(1).

In order to understand the effects of adding doubly charged ions in the presence of ECDI, Fig. 4 shows the contours of ϵr=0 and ϵi=0 as well as the roots of the 3D ECDI-IITSI dispersion relation for kyλD=0.9, which corresponds to one of the 2D ECDI resonance modes, as can be seen from Fig. 3(a2). For all cases, there are two roots that are unstable but not complex conjugates. Here, kzλD=0.01 and α=0.2 is considered.

FIG. 4.

Coupled ECDI and IITSI contours (black line: ϵr=0; red line: ϵi=0) and roots (blue circles) for various kxλD. Here, kyλD=0.9,kzλD=0.01, and α=0.2. The gray dashed lines show where γ and ωd,r=ωrkxUi+ are 0.

FIG. 4.

Coupled ECDI and IITSI contours (black line: ϵr=0; red line: ϵi=0) and roots (blue circles) for various kxλD. Here, kyλD=0.9,kzλD=0.01, and α=0.2. The gray dashed lines show where γ and ωd,r=ωrkxUi+ are 0.

Close modal

For a small kxλD that is close to 0, as shown in Fig. 4(a), the IITSI roots are close to (γ,ωd,r)=(0,0) as the origin of the doubly charged ion branch is ω̃d,r=k̃xΔŨi+0. The other root, i.e., the ECDI root, is largely unperturbed, similar to the ECDI-only case, as shown in Fig. 1(a). In Fig. 4(b), when kxλD=0.2, the singly and doubly charged ion branches separate due to the increasing kx, and the growth rate for the IITSI root starts to increase, similar to the IITSI-only case as shown in Fig. 1(b). However, the unstable root due to the ECDI mode is still dominant over the IITSI root. The presence of the IITSI lobes increases the instability growth rate of the rightmost root in Fig. 4(b) by about 0.2% compared to ECDI-only growth rate with the same k. Additionally, the rightmost ECDI root in Fig. 4(b) gets pushed toward larger real frequencies by 32% compared to the ECDI-only phase case with the same k.

For kxλD=0.4, as shown in Fig. 4(c), the ϵi=0 contour transitions from a combination of the ECDI and IITSI features into two separate ECDI type contours stemming from the singly and doubly charged ion branches. It can be seen in Fig. 4(c) that the two roots in the middle are no longer a complex conjugate due to the coupling of the ECDI and IITSI. In addition, here, the unstable root in the middle (at ω̃d,r0.4) shows a larger growth rate compared to the rightmost root (at ω̃d,r1). The growth rate of the IITSI branch becomes dominant over the ECDI root. This leads to a continuous change in γ but a discontinuous change in ωd,r as kx increases. As kxλD continues to increase through 0.8 and 1.0, as shown in Figs. 4(d) and 4(e), the shape of the contours begins to further change. Due to the separation of the singly and doubly charged ion branches as the kx increases, two separate ϵi=0 contours that are similar to the ones observed for the ECDI case [as shown in Fig. 1(a)] appear around the singly and doubly charge ion branches, while the ϵr=0 contours still remain similar to the IITSI case [as shown in Fig. 1(b)]. As the ϵi=0 contours separate, the two middle roots are not complex conjugates (e.g., the ωd,r separates and the magnitudes of γ are no longer identical). The unstable root at 0<ωd,r<ωpi has a larger growth rate than the rightmost unstable root at ωd,r>ωpi. As kxλD further increases, the central loop of ϵr=0 stretches until it eventually breaks and forms two ECDI-type contours, as shown in Fig. 4(f), i.e., kxλD=1.1. Once the ϵr=0 contours finally separate, the instability has transitioned to two separate ECDI contours. Each pair of closed contours in Fig. 4(f) is identical to the ECDI-only contour in Fig. 1(a). It can be considered in Fig. 4(f) that one closed contour is for singly charged ions, as it is centered at ωd,r=0, i.e., ωr=kxUi+, and the other closed contour is for doubly charged ions, as it is centered at ωd,r=kxΔUi+, i.e., ωr=kxUi2+.

In summary, the most unstable root due to the coupling of ECDI and IITSI can be characterized by analyzing the ϵr and ϵi contours as follows:

  • At lower kx, the growth rate of the middle (IITSI) root is smaller than that of the rightmost (ECDI) root. See Figs. 4(a) and 4(b).

  • For intermediate kx, the growth rate of the middle root grows and becomes larger than that of the rightmost root. The ϵi=0 contours separate into ECDI-like features stemming from the singly and doubly charged ion branches, while ϵr=0 remain similar to the IITSI feature, illustrating the coupling between ECDI and IITSI. See Figs. 4(c)–4(e).

  • For larger kx, two separate contours that correspond to the singly and doubly charged ions are observed. Hence, the dispersion relation is a superposition of two ECDI contours as the doubly charged ion branch ωd,r=kxΔUi+ separate from the singly charged ion branch ωd,r=0. See Fig. 4(f).

The complex root-finding algorithm discussed above is used to solve the coupled ECDI-IITSI dispersion relation for two values of α: 0.05 and 0.2. A doubly charged ion fraction up to 0.2 is found in HETs.40 Similar to Fig. 3, the growth rates and real frequencies for α=0.05 and α=0.2 cases are plotted as functions of kxλD and kyλD discretized over 180 values from 0 to 2, respectively. Here, three different kzλD values, kzλD= 0.01, 0.03, and 0.09, are shown, similar to the values used in Figs. 2 and 3(a)–3(c).

1. α = 0.05

Figure 5 illustrates the ECDI and IITSI coupling for α=0.05, which corresponds to a weak IITSI.13 In contrast to the singly charged ECDI-only case for kzλD=0.01 in Fig. 3(a) where the ECDI resonances, i.e., kyUdnωce, are sufficiently separated, Figs. 5(a2), 5(b2), and 5(c2) show that the most unstable roots are more complex, having influence from both the ECDI and IITSI. The colorbar spans the same range of values for real frequencies, as shown in Figs. 5(a1)–5(c1), and growth rates, as shown in Figs. 5(a2)–5(c2), respectively.

FIG. 5.

The drift-shifted real frequency ωd,r/ωpi (first row) and the growth rate γ/ωpi (second row) as a function of kxλD and kyλD with α=0.05: (a) kzλD=0.01, (b) kzλD=0.03, and (c) kzλD=0.09. Maximum values of the colorbar for ωd,r/ωpi=1.1 and γ/ωpi = 0.34.

FIG. 5.

The drift-shifted real frequency ωd,r/ωpi (first row) and the growth rate γ/ωpi (second row) as a function of kxλD and kyλD with α=0.05: (a) kzλD=0.01, (b) kzλD=0.03, and (c) kzλD=0.09. Maximum values of the colorbar for ωd,r/ωpi=1.1 and γ/ωpi = 0.34.

Close modal

It can be seen from the growth rate plots in Figs. 5(a2)–5(c2) that the IITSI mode stays similar for different values of kzλD. This is because the IITSI is an electrostatic instability caused by the streams of singly and doubly charged ions in the axial direction. On the other hand, the ECDI modes transition from resonance-type solution to broadband structures due to the increasing kzλD, as shown in Figs. 3(a2)–3(c2). The coexistence of the IITSI and ECDI modes results in a complex structure of the most dominant unstable solution. In the limit of the 2D case13 (i.e., kzλD0), it was found that the most unstable mode of the ECDI-IITSI solution is at kxλD=0, as the 2D ECDI growth rate due to the resonant modes [see Eq. (6)] is larger than the IITSI growth rate. This means that the plasma wave generated due to the coupled ECDI-IITSI occurs in the direction of the E×B drift in the 2D limit. Interestingly, in the 3D cases, shown in Figs. 5(a2)–5(c2), as the 3D ECDI growth rate is smaller than the 2D ECDI growth rate due to the broadband structure as kzλD increases, the maximum growth rate of the 3D ECDI-IITSI dispersion can be found to exist in the region where the IITSI occurs. The maximum growth rate is now at a finite kxλD and a finite kyλD, indicating that the plasma wave generated due to the coupled ECDI-IITSI occurs in an oblique direction (in the xy plane) for the 3D cases.

Figures 5(a1)–5(c1) show the real frequencies of the maximum growth rate, exhibiting more clearly the transition between ECDI and IITSI modes. At smaller kxλD [e.g., kxλD0.3 for Figs. 5(a) and kxλD0.15 for Figs. 5(b) and 5(c)], the maximum growth rate is large where the ECDI mode is most unstable. At a larger kxλD (e.g., kxλD0.3), the ECDI mode becomes less dominant, and the IITSI mode results in the maximum growth rate. As discussed in Figs. 4(b) and 4(c), the real frequency, ωr, of the ECDI root is larger than that of the IITSI root, which explains the discontinuity in the real frequency of the most unstable mode, as shown in Figs. 5(a1)–5(c1). At a much larger kxλD (e.g., kxλD0.8), the real frequency and the growth rate do not exhibit any discontinuity as the IITSI root of the singly and doubly charge ions, as shown in Figs. 4(c)–4(e), transitions to the ECDI root of the singly charged ion contribution, as shown in Fig. 4(f). The switching between low and high real frequency roots exemplifies the sensitivity of this coupled instability to the system dimensions, as this determine the largest allowable wavelength modes (and as such the smallest wavenumbers) in each direction. Depending on the smallest allowable wavenumber, the coupled instability can be characterized by different values for ωr.

2. α = 0.2

Figure 6 shows the real frequency and the growth rate of the most unstable solution for α=0.2, which corresponds to a strong IITSI.13 While the profile of the most unstable roots is similar to the α=0.05 case shown in Fig. 5, it can be seen from Fig. 6 that the IITSI roots become more dominant compared to the ECDI roots. The increase in the IITSI growth rate is consistent with Ref. 13. The maximum growth rate is 33% larger for the coupled α=0.2 case (Fig. 6) compared to the coupled α=0.05 case (Fig. 5). The maximum growth rate of the 3D ECDI-IITSI case with α=0.2 and kzλD=0.01 [Fig. 6(a2)] is roughly 7% larger than the IITSI-only growth rates [Fig. 3(d)]. This maximum occurs in the region near kxλD=0.7 and kyλD=1.8 where the ECDI resonance overlaps with the IITSI region of greatest instability growth. As kzλD increases, the maximum coupled ECDI-IITSI growth rate approaches that of IITSI-only.

FIG. 6.

The drift-shifted real frequency ωd,r/ωpi (first row) and the growth rate γ/ωpi (second row) as a function of kxλD and kyλD with α=0.2: (a) kzλD=0.01, (b) kzλD=0.03, and (c) kzλD=0.09. Maximum values of the colorbar for ωd,r/ωpi=1.5 and γ/ωpi = 0.36.

FIG. 6.

The drift-shifted real frequency ωd,r/ωpi (first row) and the growth rate γ/ωpi (second row) as a function of kxλD and kyλD with α=0.2: (a) kzλD=0.01, (b) kzλD=0.03, and (c) kzλD=0.09. Maximum values of the colorbar for ωd,r/ωpi=1.5 and γ/ωpi = 0.36.

Close modal

One interesting observation can be made in Fig. 6(a1). The singly charged ECDI resonance dominates the growth rate of the doubly charged ECDI resonance for many of the (kx, ky) values. However, at kxλD>1 and for large enough α, i.e., α0.2, the growth rate of the doubly charged resonance is able to surpass the growth rate of the singly charged resonance. In contrast to Fig. 5(a1), the drift-shifted frequency of the fastest growing mode in the ECDI resonances discontinuously changes as the ECDI root switches from the singly charged ion branch to the doubly charged ion branch. This can only occur for sufficiently large values of α, as the relative importance of the doubly charged ion term in the ion susceptibility is a function of kxΔUi and 2α/(1α). For α=0.2, the quantity 2α/(1α) is 0.5 compared to the α=0.05 case, in which the quantity is about 0.1, justifying why this effect is only seen in the α=0.2 case. The drift-shifted frequency of the doubly charged resonance, and hence the phase velocity of the wave, is significantly larger than the singly charged resonance. For instance, this can be seen from Fig. 4(f), where there are two unstable roots: one with ωd,r/ωpi0.8 and the other with ωd,r/ωpi2.2. In 3D, the resonances are thermally broadened, which can cause the singly charged ECDI root to have a smaller growth rate, resulting in a larger growth rate due to the doubly charged ECDI root.

One of the greatest advantages of considering the 3D ECDI and IITSI dispersion relation is the ability to capture the thermal broadening of the cyclotron resonances as kzλD increases.17 In 1D and 2D studies, the ECDI roots exhibit resonance conditions, as can be seen in Eq. (6). As a result, the growth rates of the resonances in 2D are larger than the 3D case for a given electric field, magnetic field, electron temperature, and electron density. In the 2D ECDI-IITSI dispersion, i.e., k=0,13 the ECDI growth rates dominate the IITSI growth rates for all values of kxλD, i.e., the largest growth rates occur at kxλD=0. This means the plasma wave initiated by 2D ECDI-IITSI traverses in the azimuthal direction with the allowable ky values within the plasma device of interest, e.g., Hall effect thruster discharge channel. However, in the 3D ECDI-IITSI case (i.e., kz0), the maximum growth rate of the ECDI resonances becomes lower and widened, which resembles a transition from the resonant modes to broadband ion-acoustic modes, as kz increases. Consequently, the IITSI growth rates can come to dominate over the ECDI growth rates for intermediate kxλD. For kzλD=0.01, the maximum growth rate occurs in the IITSI region along the ECDI resonances, as the ECDI and IITSI couple. The first three ECDI resonances at kyλD0.5,0.9, and 1.4 for α=0.05 have the largest growth rates with wavenumbers at kxλD0.64, corresponding to waves propagating at 48°, 34°, and 27° with respect to the axial electric field, respectively. Thus, if ECDI and IITSI are present in the same region within HETs, the results suggest that the resulting waves driven by the coupled ECDI-IITSI propagate at some oblique angle to the axial electric field, rather than purely in the azimuthal direction as previously predicted by two-dimensional theory. It is important to note that finite ion temperature effects could serve to stabilize the coupled instability,28 which is reserved for future work. Furthermore, because this prediction is made with linear dispersion theory, the exact values of the propagation angle with respect to the axial direction may be altered due to nonlinear effects. Nevertheless, considering the 3D kinetic dispersion relation for coupled ECDI and IITSI may explain the experimental evidence of axially propagating modes observed by Tsikata et al.12 

The analysis of dispersion relations using complex root finding enables the study of plasma kinetic instabilities that have multiple unstable roots. The algorithm for solving dispersion relations discussed in this paper can be extended to any arbitrary dispersion relation that has more than one unstable root to evaluate the most unstable mode. As long as the dispersion relation can be numerically evaluated for various ω=ωr+iγ values, the intersections of the contours of Re(ϵ(ω,k))=0 and Im(ϵ(ω,k))=0 can be found, yielding valid solutions to the dispersion relation. While not as computationally efficient as the fixed-point algorithm of Cavalier et al.,17 this complex root-finding algorithm is a robust method for analyzing complicated dispersion relations.

Applying the complex root-finding algorithm to the three-dimensional coupling of ECDI and IITSI reveals interesting behavior in ωr and γ as the wavenumber and direction of propagation changes, including the switching between ECDI and IITSI roots for intermediate values of wavenumber parallel to the electric field. The largest difference between the two-dimensional and three-dimensional study of the coupling of ECDI and IITSI is the location of the largest growth rates. In previous two-dimensional studies,13 the ECDI growth rate at kx = 0 always dominates the IITSI growth rates at finite kx. However, the present three-dimensional theory indicates that the highest growth rate for a given ky occurs at finite kx, resulting in a wave that propagates at an oblique angle to the electric field. This finding gives new insight into possible waves present in the plume regions of Hall effect thrusters in which both ECDI and IITSI are present.

The authors acknowledge the discussions with S. Tsikata, J. C. Adam, and A. Héron. This work was supported by a NASA Space Technology Graduate Research Opportunity, under Grant No. 80NSSC21K1273, NASA through the Joint Advanced Propulsion Institute, a NASA Space Technology Research Institute under Grant No. 80NSSC21K1118, and the U.S. Department of Energy, Office of Science, Office of Fusion. Energy Sciences, under Award No. DE-SC0020623.

The authors have no conflicts to disclose.

Andrew Christopher Denig: Conceptualization (equal); Data curation (lead); Formal analysis (lead); Methodology (lead); Visualization (equal); Writing – original draft (lead); Writing – review & editing (equal). Kentaro Hara: Conceptualization (equal); Formal analysis (supporting); Funding acquisition (lead); Methodology (supporting); Supervision (lead); Visualization (equal); Writing – original draft (supporting); Writing – review & editing (equal).

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

1.
D. W.
Forslund
,
R. L.
Morse
, and
C. W.
Nielson
,
Phys. Rev. Lett.
25
(
18
),
1266
1270
(
1970
).
2.
D. W.
Forslund
,
R. L.
Morse
, and
C. W.
Nielson
,
Phys. Rev. Lett.
27
(
21
),
1424
1428
(
1971
).
3.
J. B.
McBride
,
E.
Ott
,
J. P.
Boris
, and
J. H.
Orens
,
Phys. Fluids
15
,
2367
2383
(
1972
).
4.
C.
Lashmore-Davies
and
T.
Martin
,
Nucl. Fusion
13
,
193
203
(
1973
).
5.
S.
Tsikata
,
N.
Lemoine
,
V.
Pisarev
, and
D. M.
Grésillon
,
Phys. Plasmas
16
,
033506
(
2009
).
6.
S.
Tsikata
and
T.
Minea
,
Phys. Rev. Lett.
114
(
18
),
185001
(
2015
).
7.
T.
Lafleur
,
S. D.
Baalrud
, and
P.
Chabert
,
Phys. Plasmas
23
,
053503
(
2016
).
8.
J. P.
Boeuf
and
L.
Garrigues
,
Phys. Plasmas
25
,
061204
(
2018
).
9.
S.
Janhunen
,
A.
Smolyakov
,
D.
Sydorenko
,
M.
Jimenez
,
I.
Kaganovich
, and
Y.
Raitses
,
Phys. Plasmas
25
,
082308
(
2018
).
10.
T.
Charoy
,
J. P.
Boeuf
,
A.
Bourdon
,
J. A.
Carlsson
,
P.
Chabert
,
B.
Cuenot
,
D.
Eremin
,
L.
Garrigues
,
K.
Hara
,
I. D.
Kaganovich
,
A. T.
Powis
,
A.
Smolyakov
,
D.
Sydorenko
,
A.
Tavant
,
O.
Vermorel
, and
W.
Villafana
,
Plasma Sources Sci. Technol.
28
,
105010
(
2019
).
11.
W.
Villafana
,
F.
Petronio
,
A. C.
Denig
,
M. J.
Jimenez
,
D.
Eremin
,
L.
Garrigues
,
F.
Taccogna
,
A.
Alvarez-Laguna
,
J. P.
Boeuf
,
A.
Bourdon
,
P.
Chabert
,
T.
Charoy
,
B.
Cuenot
,
K.
Hara
,
F.
Pechereau
,
A.
Smolyakov
,
D.
Sydorenko
,
A.
Tavant
, and
O.
Vermorel
,
Plasma Sources Sci. Technol.
30
,
075002
(
2021
).
12.
S.
Tsikata
,
J.
Cavalier
,
A.
Héron
,
C.
Honoré
,
N.
Lemoine
,
D.
Grésillon
, and
D.
Coulette
,
Phys. Plasmas
21
,
072116
(
2014
).
13.
K.
Hara
and
S.
Tsikata
,
Phys. Rev. E
102
(
2
),
023202
(
2020
).
14.
P.
Kumar
,
S.
Tsikata
, and
K.
Hara
,
J. Appl. Phys.
130
,
173307
(
2021
).
15.
S. T.
Sewell
,
P.
Kumar
, and
K.
H
,
IEEE Trans. Plasma Sci.
50
(
10
),
3498
3506
(
2022
).
16.
F.
Chen
,
Introduction to Plasma Physics and Controlled Fusion Introduction to Plasma Physics and Controlled Fusion
(
Springer
,
1984
).
17.
J.
Cavalier
,
N.
Lemoine
,
G.
Bonhomme
,
S.
Tsikata
,
C.
Honoré
, and
D.
Grésillon
,
Phys. Plasmas
20
,
082107
(
2013
).
18.
J. P.
Boeuf
,
G.
Fubiani
, and
L.
Garrigues
,
Plasma Sources Sci. Technol.
25
,
045010
(
2016
).
19.
K.
Hara
,
Y.
Yamashita
,
S.
Tsikata
,
B.
Vincent
,
S.
Mazouffre
, and
S.
Cho
, “
New insights into electron transport due to azimuthal drift in a Hall effect thruster
,” IEPC-2019-691, in
36th International Electric Propulsion Conference
,
Vienna, Austria
,
2019
.
20.
M. D.
Campanell
,
A. V.
Khrabrov
, and
I. D.
Kaganovich
,
Phys. Rev. Lett.
108
(
23
),
235001
(
2012
).
21.
D.
Sydorenko
,
A.
Smolyakov
,
I.
Kaganovich
, and
Y.
Raitses
,
Phys. Plasmas
14
,
013508
(
2007
).
22.
J. P.
Boeuf
, “
Cross-field diffusion in Hall thrusters and other plasma thrusters
,” in
APS Annual Gaseous Electronics Meeting Abstracts
,
2012
.
23.
I. G.
Mikellides
and
I.
Katz
,
Phys. Rev. E
86
(
4
),
046703
(
2012
).
24.
Y.
Sakawa
,
C.
Joshi
,
P. K.
Kaw
,
F. F.
Chen
, and
V. K.
Jain
,
Phys. Fluids B
5
,
1681
1694
(
1993
).
25.
A. I.
Smolyakov
,
O.
Chapurin
,
W.
Frias
,
O.
Koshkarov
,
I.
Romadanov
,
T.
Tang
,
M.
Umansky
,
Y.
Raitses
,
I. D.
Kaganovich
, and
V. P.
Lakhin
,
Plasma Phys. Controlled Fusion
59
,
014041
(
2016
).
26.
K.
Hara
,
A. R.
Mansour
, and
S.
Tsikata
,
J. Plasma Phys.
88
,
905880408
(
2022
).
27.
Y.
Nariyuki
and
T.
Hada
,
J. Geophys. Res.: Space Phys.
112
, A10107, (
2007
).
28.
J.
Cavalier
, “
Modeles Cinetiques et Caracterisation Experimentale Des Fluctuations Electrostatiques Dans un Propulseur a Effet Hall
,” Ph.D. thesis (
Universite de Lorraine
,
2013
).
29.
S. D.
Baalrud
and
J.
Daligault
,
AIP Conf. Proc.
1786
,
130001
(
2016
).
30.
K.
Hara
and
S.
Cho
, “Radial-azimuthal particle-in-cell simulation of a Hall effect thruster,” IEPC-2017-495, in 35th International Electric Propulsion Conference, Atlanta, GA, 2017.
31.
A.
Ducrocq
,
J. C.
Adam
,
A.
Héron
, and
G.
Laval
,
Phys. Plasmas
13
,
102111
(
2006
).
32.
I. B.
Bernstein
,
Phys. Rev.
109
(
1
),
10
21
(
1958
).
33.
S.
Tsikata
, “
Small-scale electron density fluctuations in the Hall thruster investigated by collective light scattering
,” Ph.D. thesis (
Ecole Polytechnique
,
2009
).
34.
G.
Gordeev
,
JETP
6
,
660
(
1952
).
35.
W.
Gautschi
,
SIAM J. Numer. Anal.
7
,
187
198
(
1970
).
36.
M. R.
Zaghloul
and
A. N.
Ali
,
ACM Trans. Math. Software
38
,
1
–22 (
2012
).
37.
S. D.
Baalrud
,
Phys. Plasmas
20
,
012118
(
2013
).
38.
D.
Schwarz
, see https://www.mathworks.com/matlabcentral/fileexchange/11837-fast-and-robust-curve-intersections for “Fast and robust curve intersection,” MATLAB Central File Exchange; accessed 30 June
2021
.
39.
S. P.
Gary
and
J. J.
Sanderson
,
J. Plasma Phys.
4
,
739
751
(
1970
).
40.
R. R.
Hofer
and
A. D.
Gallimore
,
J. Propul. Power
22
,
732
740
(
2006
).