Low-temperature plasmas are used in various applications, such as Hall thrusters for satellite propulsion, ion sources and magnetron discharges for plasma processing, and negative ion sources for neutral beam injection in fusion. The plasmas in these devices are partially magnetized, meaning that the electrons are strongly magnetized while the ions are not. They are subject to various micro- and macro-instabilities that differ significantly from instabilities in fusion plasmas. These instabilities are often triggered by the large difference in electron and ion drift velocities in the direction. The possibility of maintaining a large electric field in the quasineutral plasma of Hall thrusters despite anomalous electron transport, or the presence of strong double layers associated with the azimuthal rotation of plasma structures (“rotating spokes”) in magnetron discharges and Hall thrusters are examples of the very challenging and exciting physics of devices. The turbulence and instabilities present in plasma devices constitute a major obstacle to the quantitative description of these devices and to the development of predictive codes and are the subject of intense research efforts. In this tutorial, we discuss the key aspects of the physics of low-temperature partially magnetized plasmas, as well as recent advances made through simulations, theory, and experiments in our understanding of the various types of instabilities (such as gradient-drift/Simon-Hoh and lower hybrid instabilities, rotating ionization waves, electron cyclotron drift instability, modified two-stream instability, etc.) that occur in these plasmas.
I. INTRODUCTION
In low-temperature plasmas devices, an external magnetic field is placed perpendicular to the discharge current or to the electric field applied between two electrodes. The perpendicular magnetic field increases the electron residence time in the device and the electron collisionality, allowing plasma sustainment by electron impact ionization at low pressures and moderate applied voltage. The perpendicular magnetic field decreases the electron conductivity parallel to the electric field, so an objective of the configuration can also be to generate a large electric field in the quasineutral plasma, which can accelerate ions (application to gridless ion sources such as the Hall thruster). Lowering the electron conductivity with a perpendicular magnetic field is actually not trivial because of the development of instabilities that tend to increase electron transport across the magnetic field. This question is at the center of this tutorial.
In most applications of plasmas, the external magnetic field and the device dimensions are such that electrons are strongly magnetized (Larmor radius much smaller than the device dimensions), while ions are not or are only weakly magnetized. Such plasmas are called “partially magnetized plasmas.” Electron transport along the electric field is, in principle, strongly reduced by the external magnetic field applied perpendicular to the electric field. In most devices, the magnetic field lines intersect a wall (cathode or dielectric surface) so that along the magnetic field line the electrons are electrostatically confined by the wall sheath potential. Electron confinement is therefore a combination of magnetic and electrostatic confinement. Collisionless electrons are trapped in the direction perpendicular to the magnetic field. Only collisions with neutrals can un-trap the electrons from their cyclotron trajectory and allow transport across the magnetic field (“cross field transport”) along the electric field. Electron confinement by the magnetic field in the direction of the applied electric field is generally not as good as predicted by the classical collisional theory because of turbulence and instabilities that lead to cross field transport (“anomalous transport”). Because of their partial magnetization, the physics and instabilities of these plasmas are very distinct from those of other plasmas, e.g., for fusion applications.
The study of instabilities and anomalous transport in plasmas has recently attracted much attention, largely driven by the need to better understand the operation of devices, such as Hall thrusters and magnetron discharges, and to develop predictive models of these devices. Joint efforts by the low-temperature plasma community and the plasma propulsion community “to develop a rigorous understanding of the rich phenomena observed in devices” are described in a recent perspective article by Kaganovich et al.1 that “presents the current understanding of the physics of these phenomena and state-of the-art computational results, identifies critical questions, and suggests directions for future research”. The present tutorial article can be considered as complementary to the perspective article by Kaganovich et al., but is aimed at non-specialists, with more emphasis on the physics of instabilities in various plasma devices.
In plasmas with perpendicular electric field and magnetic field, there is a charged particle drift in the direction (“ drift”) with a velocity close to . This drift plays an important role in regions of large electric fields. Electrons are subject to this drift while ions are accelerated out of the large electric field region before they have time to reach the drift velocity (this is another way to say that the plasma is partially magnetized). The large electron drift and the negligible ion drift in this direction can lead to charge separation and is often responsible for triggering instabilities. The electron pressure gradient perpendicular to the magnetic field also creates a drift in the direction. This is the diamagnetic drift velocity, where is the plasma density, the elementary charge, and the electron temperature in eV. and electron drift can both contribute to the development of instabilities. Instabilities are ubiquitous in low-temperature partially magnetized plasmas and can significantly enhance cross field electron transport. They manifest themselves in a wide range of unstable fluctuations, with frequencies from kilohertz to tens of megahertz and wavelengths from a fraction of millimeter to several centimeters.2
The electron drift in the direction gives rise to the Hall current. Efficient electron confinement parallel to the electric field requires that the Hall current does not intercept a wall (otherwise the Hall effect would destroy the confinement of electrons). This is possible if the Hall current is closed on itself, as for example, in a cylindrical geometry where the electric field is axial and the magnetic field radial, or the opposite. The Hall current in such geometries is in the azimuthal direction, i.e., is closed on itself. Plasma devices with closed electron drift are called closed-drift devices.3
The concept of closed-drift devices has led to important applications in ion sources for space propulsion (Hall thrusters, HT, also called stationary plasma thrusters, SPT,4–7 and thrusters with anode layer, TAL5) and plasma processing (End-Hall ion sources,3,8,9 closed-drift anode layer ion sources, ALIS,3,10 or Anode Layer Plasma Accelerators, ALPA3,11) The originality of these “Hall ion sources” lies in the fact that ions are extracted from the plasma not by a system of biased grids (they are “gridless ion sources”), but by the electric field generated in the quasineutral plasma by the local lowering of electron conductivity due to a magnetic field barrier placed between an emissive cathode and the anode. Magnetron discharges can also be considered as closed-drift ion sources where ions are extracted from the plasma and accelerated in the cathode sheath to sputter atoms from the cathode, with applications to thin film deposition. They can be operated at high power densities in a pulsed regime (HiPIMS, high power impulse magnetron sputtering).12–15 In magnetron discharges as in Hall thrusters, electrons are heated by the applied dc electric field. They drift in the azimuthal direction, generate the plasma by ionization and slowly drift to the anode due to collisions or instabilities. Penning discharges,16–19 consisting of two facing cathodes surrounded by a cylindrical anode, with an axial magnetic field, are another type of closed-drift device that has been used in many ion source applications. Hall ion sources, magnetron discharges, and Penning ion sources operate under pressure typically between a fraction of pascals and a few pascals but it is interesting to note that Penning discharges where first implemented as ion gauges20,21 in the 1930s, operating at much lower pressures. In these conditions, the plasma can become non-neutral (electron density larger than ion density) and specific types of instabilities, called diocotron instabilities22–26 are likely to develop.
Historical references on the development of devices for electric propulsion can be found in Refs. 3, 4, and 27, for magnetron discharges applications to sputtering and deposition in Refs. 12–15, and for various cross field devices operating over a wide pressure range in Refs. 23 and 28. Recent review or perspective papers on electric propulsion including gridless ion sources can be found in Refs. 29–31.
The question of cross field electron transport or anomalous transport also arises in situations where the plasma simply diffuses across the magnetic field, in the absence of applied voltage perpendicular to the magnetic field. The plasma can be generated by inductive coupling, microwave, of by injection of an electron beam. This is the case of magnetized plasma sources where the plasma is confined by magnetic cusps (applications to plasma processing or gridded ion sources32–34) or of magnetized cylindrical plasma columns with axial magnetic field. Magnetized plasma columns have been used extensively to study turbulence and anomalous electron transport.35–43, and flows combine in these plasmas to ensure a zero net current to the walls. Diffusion is often non-ambipolar and “short-circuit” effects44 are possible in the case of metallic walls, i.e., the electron current does not balance locally but globally the ion current and a large fraction of the electron current to the walls is in the parallel direction (along the magnetic field). plasmas have potential applications in mass separation, which requires the imposition and customization of an electric field perpendicular to the axial magnetic field in a plasma column to create an flow.45,46 Research is ongoing into ways to generate this radial electric field, such as through electrode biasing,46–48 the use of emissive end-electrodes,49 or wave injection.50
plasma devices, such as Hall ion sources or magnetron discharges, are now currently used successfully in space propulsion and plasmas processing applications. However, despite important research efforts and progress in the last two or three decades, the physics of these devices is still poorly understood and models of these plasmas with predictive capabilities are still lacking. In this article we describe, for non-specialists, the basic physics of low-temperature partially magnetized plasmas, with emphasis on the development of instabilities, and we illustrate with some examples the recent progress in our understanding of these plasmas and their instabilities through experiments, simulation, and theory.
The basic physics and principles of plasma devices are presented in Sec. II. Section III shows how instabilities in plasmas can be described and understood with the combination of theory (dispersion relations) and simulations. Section IV focuses on two important plasma devices, the Hall thruster and the magnetron discharge, with a more detailed discussion of the physics and of the evidence of the presence of instabilities in these devices. Microscopic (kinetic) instabilities in Hall thrusters are studied in Sec. V with the help of a kinetic dispersion relation and particle simulations. The macroscopic instabilities that can be observed in magnetron discharges and under certain Hall thruster conditions (rotating spokes) are described in Sec. VI with particle simulations and comparisons with experiments.
II. BASIC PHYSICS AND PRINCIPLES OF E × B PLASMA DEVICES
The typical conditions of low-temperature partially magnetized plasmas devices and the orders of magnitude of the plasma parameters are presented in Subsection II A. The classical, collisional transport theory is briefly recalled in Subsection II B. The geometry and principles of the main closed-drift devices are described in Subsection II C. The Hall effect, which occurs in devices with non-closed-drift, is discussed with two examples in Subsection II D.
A. Conditions and orders of magnitude
Under usual conditions of plasmas applications, the effect of the plasma current on the magnetic field is negligible so we will assume that the magnetic field is purely external, i.e., is generated by a system of coils or magnets and a magnetic circuit. This condition may not be satisfied51 in the extreme conditions of HiPIMS where very high power densities are injected in the plasma through short pulses of several microseconds.
As said in the introduction, we consider partially magnetized plasmas, which implies that the mean electron Larmor radius is much smaller than the device dimensions while the ion Larmor radius is not. For a Maxwellian electron velocity distribution at temperature , the mean Larmor radius is given by where is the electron mean velocity perpendicular to the magnetic field, and is the electron cyclotron angular frequency. Orders of magnitude of magnetic field intensity, charged particles cyclotron frequencies, and Larmor radii are given in Table I. Electron–neutral collisions play an important role in applications of plasmas because they control the electron magnetization and contribute to cross field transport (and also because the plasma is sustained by electron impact ionization). Hall ion sources and magnetron discharges operate at pressures typically between 0.1 and 5 Pa and with magnetic field usually smaller than 100 mT. The plasma dimensions are generally less than 5–10 cm. Table II gives some orders of magnitude of collision frequencies, and free paths of electron and argon ions.
(mT) . | . |
---|---|
(rad/s) | |
(μm) for 5 eV electrons | |
(rad/s) Ar+ ions | |
(cm) for 5 eV Ar+ ions |
(mT) . | . |
---|---|
(rad/s) | |
(μm) for 5 eV electrons | |
(rad/s) Ar+ ions | |
(cm) for 5 eV Ar+ ions |
(Pa) − 300 K . | . |
---|---|
(mtorr) | |
(m−3) | |
(s−1) for 5 eV electron | |
(s−1) for 5 eV Ar+ ions | |
(cm) for 5 eV electron | |
(cm) for 5 eV Ar+ ions |
(Pa) − 300 K . | . |
---|---|
(mtorr) | |
(m−3) | |
(s−1) for 5 eV electron | |
(s−1) for 5 eV Ar+ ions | |
(cm) for 5 eV electron | |
(cm) for 5 eV Ar+ ions |
Usual values of plasma densities, plasma frequencies, and electron Debye lengths are given in Table III.
(m−3) . | . |
---|---|
(s−1) | |
(s−1) for Ar+ ions | |
(μm) for 5 eV |
(m−3) . | . |
---|---|
(s−1) | |
(s−1) for Ar+ ions | |
(μm) for 5 eV |
Since the electron trajectory around the magnetic field changes after each electron–neutral collision, leading to transport or diffusion across the magnetic field, an important parameter is the ratio of the electron Larmor frequency over the electron–neutral collision frequency. This ratio, , is called the Hall parameter and characterizes the electron confinement by the magnetic field. As said in the introduction, plasma fluctuations and instabilities also contribute to un-trapping the electrons from their trajectory around the magnetic field. Their effect can be considered as similar to the effect of collisions, and one can define an effective Hall parameter and an effective collision frequency to describe their effect on cross field transport. It can be deduced from Tables I and II that the Hall parameter associated with electron–neutral collisions, for a magnetic field intensity of 10 mT, and a gas pressure of 0.1 Pa is .
Note finally that low-temperature partially magnetized plasmas are far from thermal equilibrium, and that the electron temperature (defined as of the thermal energy) is much larger than the ion temperature and the charged particle velocity distribution functions can be very far from Maxwellian.
B. Classical, collisional cross field transport
In this section, the classical, collisional expression of the electron current across a magnetic field is derived from a simplified form of the steady-state, electron momentum equation. We neglect the inertia term and assume that the electron pressure tensor is scalar. We consider a plasma with an electric field is in the direction, the magnetic field in the perpendicular direction. We assume that the plasma density gradient is parallel to the electric field, in the direction (as in Hall thrusters or magnetron discharges). The notations are indicated in Fig. 1.
Along the magnetic field lines (in the direction), electrons move freely between collisions. The electron conductivity parallel to the magnetic field is, therefore, much larger than the electron conductivity parallel to the electric field. An important consequence is that the magnetic field lines tend to be equipotential. More precisely, from the momentum equation along the magnetic field lines and assuming negligible electron current parallel to the magnetic field, the electric field along these lines must balance the electron pressure gradient. The potential drop along the magnetic field lines is therefore small for low electron temperature or small electron pressure gradient. Note that a magnetic field line should never connect cathode and anode in a magnetized plasma device since this would be equivalent to a short circuit at low pressure.
If the plasma is bounded by walls in the , direction, e.g., dielectric walls, there is an electron flow to one of the walls due to the and drifts. The electron flow to the wall must be balanced by an ion flow. An electric field is, therefore, generated in the direction to draw the ions to the wall. This is the Hall effect. This electric field, combined with the magnetic field, will in turn create an electron drift in the direction that will destroy the electron confinement. The question of Hall effect in low-temperature partially magnetized plasmas has been illustrated analytically in the case of a plasma column in a transverse magnetic field by Ecker and Kanne52 and by Kunkel,53 and with particle simulations54 in the case of the negative ion source for the neutral beam injector of ITER. This is summarized in Subsection II D. The solution to avoid the Hall effect is to choose a geometry where the electron current is closed on itself. This is possible in a cylindrical geometry where the electric field is axial and the magnetic field is radial or the opposite. Devices with this configuration are called closed-drift devices3 and are described in Subsection II C.
C. Closed drift devices
Some classical examples of closed-drift devices are displayed in Fig. 2. In Penning cells and cylindrical magnetrons discharges [Figs. 2(a) and 2(b)], the external magnetic field is axial, and the applied electric field is radial. The concept of Penning discharge [Fig. 2(a)] was introduced by Phillips at the end of the 19th century55 and was used for the first time in 1937 by Penning, with the construction of an ionization gauge.20 The different modes of a Penning discharge with a long anode cylinder at low pressure were studied by Knauer56 and Schuurmann16,17 and the diocotron instability that develops in Penning discharge at low pressures was described by Knauer.22 More recent experimental work on Penning discharges and their different regimes can be found in the article by Mamedov et al.57
Penning discharges are currently used as ion sources19 (ions can be extracted through an aperture in one of the two cathodes or through the anode) in various applications, e.g., to sputtering and deposition, ion implantation,58 isotope separation, production of hydrogen negative ions,59 and neutron generators.60 The electron vortices associated with the diocotron instability, which develop in Penning discharges and in cylindrical magnetrons at very low pressures, were extensively studied by Kervalishvili and his group25,61–64 with the help of probe measurements and described with particle simulations in Ref. 26. Cylindrical magnetrons [Fig. 2(b)] consist of two concentric cylindrical electrodes with an axial magnetic field. Early sputtering-deposition applications were demonstrated in cylindrical magnetrons,14,65 but the planar magnetron configuration [Fig. 2(c)] is now much more commonly used for deposition of metallic or compound thin films in many industrial applications. The planar magnetron can operate in a dc regime in the pressure range 0.2–4 Pa with applied voltages of 300–700 V and a magnetic field of 20–50 mT close to the cathode.14 Very high power densities can be achieved in a pulsed regime (HiPIMS)13,15 with repetition frequency in the range 50–5000 Hz and 1%–3% duty cycle,12–14 applied voltage in the range 500–2000 V and with peak power densities up to several kW/cm2. The plasma density is – m−3 in the dc case and can be larger than m−3 in the pulsed case. The physics of HiPIMS is extremely complex because of the high plasma density, ionization of the sputtered atoms, instabilities, neutral depletion, etc.
The magnetic field in planar magnetrons is generated by magnets placed behind the cathode as schematically shown in Fig. 2(c). In a Hall ion source [Fig. 2(d)], the objective is to generate an electric field in the quasineutral plasma and to use this field to extract ions from the plasma. The basic idea is, therefore, to place a magnetic barrier perpendicular to the applied electric field between an emissive cathode and the anode, as shown schematically in Fig. 2(d).
Figure 3 shows some practical examples of Hall ion sources. In a Hall thruster5,7,66 [Fig. 3(a)], the plasma is formed in a channel between two concentric cylinders. The anode is at one end of the channel, and an electron emissive cathode is placed outside the channel. A radial magnetic field with a maximum at the other end of the channel (“exhaust plane”) is generated by coils or magnets and a magnetic circuit. This magnetic field forms a magnetic barrier between cathode and anode. The applied electric field tends to be larger in the region of maximum magnetic field because of the lowering of the electron conductivity due to the magnetic barrier. Ions generated in the channel are extracted by this electric field. The external emissive cathode provides electrons to sustain the discharge as well as to neutralize the extracted ion beam. Several variations around this concept have been proposed and developed. Some of these designs are described in recent review papers.29,30,67
Here, we only give a brief description of some of these designs to illustrate the different possible configurations. The configuration of a magnetically shielded Hall thruster68–70 is shown in Fig. 3(b). The magnetic barrier in this thruster is such that the magnetic field in the channel, close to the exhaust plane is parallel to the dielectric walls. Since the magnetic field lines tend to be equipotential (because of the large electron conductivity along these lines), this configuration limits the electron and ion fluxes to the walls and reduces ion erosion in this region. The cylindrical Hall thruster71–73 [Fig. 3(c)], where the inner cylinder is shorter, is more adapted to small or low-power thrusters since, according to the empirical scaling laws of Hall thrusters, the channel width of the channel between the two dielectric cylinder should decrease with decreasing power. The magnetic field barrier in this thruster is not perpendicular to the thruster axis so the electric field is not perfectly axial, and the divergence of the ion beam is larger than in standard Hall thrusters. The anode layer thruster (or thruster with anode layer, TAL)74,75 is shown in [Fig. 3(d)]. In this thruster, the anode is very close to the magnetic barrier and the dielectric walls are no longer needed and are replaced by metallic walls. Ions are accelerated in a thin region close to the anode. Ion sources based on the same concept as TAL are also used in plasma processing applications under the name of ALIS3,10 (anode layer ion sources) or ALPA11,76 (Anode Layer Plasma Accelerators). Another thruster configuration where the anode is in the magnetic barrier is the wall-less thruster77 of [Fig. 3(e)] proposed by Kapulkin et al.78 in the 1990s, and by Mazouffre et al.79–81 later. A variation of a wall-less thruster is based on a standard Hall thruster with a short channel but with a ring anode placed at the channel exit. The ion acceleration in this design takes place outside the channel. The advantage of the configuration of Fig. 3(e) is its simplicity and reduced erosion problems, but the divergence of the ion beam can be an issue. Finally, gridless ion sources are not only used for spacecraft propulsion but also for plasma processing applications. This is the case of the ALIS and ALPA source mentioned above and of the End-Hall ion sources10 of Fig. 3(f), with a truncated conical anode and a magnetic field barrier along the anode surface. This ion source is used for ion assisted thin film deposition.82 Particle simulations of End-Hall ion sources can be found in Refs. 9 and 83.
A final remark on the magnetic field configuration of closed-drift devices is that in all cases, the magnetic field lines intersect the cathode or some dielectric surfaces. Electrons move freely along the magnetic field lines, so, as said in the introduction, the electron confinement is not purely magnetic but is a combination of magnetic confinement in the direction perpendicular to the applied electric field and electrostatic confinement along the magnetic field lines. In configurations where the magnetic field lines end on the cathode (Penning or magnetron discharges), plasma electrons are efficiently confined electrostatically because of the large potential drop in the cathode sheaths. In configurations where the magnetic field lines intersect a dielectric surface (most of the Hall ion sources), low-energy electrons are electrostatically confined by the wall sheath potential, but higher energy electrons can overcome this wall potential and reach the wall. Electron–wall interactions and the resulting secondary electron emission can play an important role in these devices.1
From this brief review, it appears that in closed-drift devices with axial electric field and radial magnetic field (magnetron discharges and Hall ion sources), there can be three generic types of axial profiles of the radial magnetic field, as shown in Fig. 4. The three profiles correspond to (1), magnetron discharges, with a magnetic field decreasing from cathode to anode, (2) standard Hall thrusters, with a magnetic barrier between cathode and anode, and (3) ion sources with anode layers, where the magnetic field increases from cathode to anode. The choice between these three configurations depends on the application and on the objective to be achieved (performance, lifetime, simplicity,…). We will see in the rest of this paper that the magnetic field profile has also important consequences on the nature of the instabilities that develop in these devices and on their consequences on anomalous electron transport.
D. E × B plasmas with non-closed drift: The Hall effect
The Hall effect occurs when the current is not closed on itself but flows to a wall. There is no Hall effect in closed-drift devices like the Hall thruster even though Hall thrusters are sometimes called “Hall effect thrusters.” To illustrate the Hall effect, we consider, as in Kunkel53 (see also Ecker52 for the case of a cylindrical column) a plasma slab between two dielectric or floating walls parallel to the direction and perpendicular to the direction, an electric field in the direction and a perpendicular magnetic field in the direction. The electric field must adjust so that charged particle losses to the walls are balanced by ionization, as in a positive column plasma. The electron current, in the direction, flows to the wall. These conditions are represented schematically in Fig. 5.
The mean velocity parallel to the electric field , averaged in the direction is . This is to be compared to the mean velocity in the absence of Hall effect, given by Eq. (2), . Since the ratio of ion to electron mobility without magnetic field, can be on the order of 10–2 or less, we have generally . The electron mobility parallel to the electric field in the presence of Hall effect is therefore considerably larger than without Hall effect and, in the limit , , and we have . We can conclude that the Hall effect tends to destroy the electron confinement perpendicular to the magnetic field.
The simple example above corresponds to an axially uniform plasma column between parallel walls in the presence of a perpendicular magnetic field.
We consider now the more complex example of a negative ion source used for the neutral beam injection system of ITER.84 The design of this source is schematically represented in Fig. 6. The source consists of a driver where electrons are heated by inductive coupling, a diffusion region with a magnetic filter, and an extraction region. The magnetic field is perpendicular to the plane of Fig. 6. The plasma is produced in hydrogen or deuterium at 0.3 Pa. The negative ions of hydrogen or deuterium are generated in the plasma volume by attachment on vibrationally excited states, and on the surface of the plasma grid covered with a cesium layer. Hydrogen or deuterium atoms hitting the caesiated grid are converted into negative ions. Negative ions are extracted through a system of biased grids. Negative ions can be destroyed by fast electrons coming from the driver. The objective of the magnetic filter in the diffusion region is to increase the residence time of electrons and increase their energy losses through collisions with neutrals. A second objective is also to lower the density of electrons close to the extraction region and to limit their extraction through the grids (in addition to the magnetic filter, a magnetic field is also generated by magnets near and parallel to each grid hole to limit this flux).
The plasma density gradient from the driver to the magnetic filter due to the large electron pressure in the driver creates a charged particle flux toward the extraction grids. The grid voltage is positive with respect to the grounded walls to enhance the flux of negative ions to the extraction system. In this configuration, there is an axial flux of charged particles perpendicular to the magnetic field. We therefore expect to have a combination of diamagnetic drift and drift in the direction of Fig. 6. These drifts are not closed and are perpendicular to the rectangular chamber walls, so we expect a Hall effect to occur. This Hall effect is discussed in Refs. 85–88, on the basis of PIC-MCC simulations. We present here a few results from Ref. 85. Figure 7 shows the 2D distributions of the electron density and plasma potential in the ion source. The plasma potential in the driver is much larger than in the diffusion chamber because electrons are heated in the driver and lose energy in the filter.
We observe on the potential and plasma density distributions an asymmetry, which can be attributed to the Hall effect. There is a potential difference (Hall voltage) between the sheath edges on the top and bottom walls. The plasma density is higher on the lower side of the chamber (while the electron temperature is lower on the lower side, see Ref. 89).
Figure 8 displays the intensity and streamlines of the electron and ion current densities in the chamber. We can clearly see the combination of the different electron drifts in the direction at the driver exit (diamagnetic and drift; a drift due to the magnetic field gradient, proportional to is also present). The more important plasma asymmetry predicted by the model at the driver exit has been confirmed by experiments.90 Compared with a closed-drift system, the Hall effect enhances the electron flow through the magnetic filter, and no instability is observed in these conditions.
If the walls are removed in the direction and periodic conditions are used instead (see Ref. 91), the drift is closed on itself and electron confinement in the axial direction is much more efficient. In these conditions, instabilities are present,91 which is not the case in the bounded conditions of Figs. 7 and 8. Finally, we note, by comparing Figs. 8(a) and 8(b), that the plasma diffusion in this source is clearly non-ambipolar except in the magnetic field free driver. Non-magnetized ions simply go down the potential while the streamlines of the electron current density are much more complex. This feature is typical of partially magnetized plasmas.
III. THEORY AND MODELING OF INSTABILITIES IN E × B PLASMAS
The basic way to study the existence of instabilities in a physical system is to look at how perturbations around a given steady state grow in time or are damped. This is done by linearizing the equations characterizing the steady state regime. In plasmas, instabilities are described by linearizing transport equation coupled with field equations. For some instabilities, a fluid approach of the transport equations is sufficient (macroscopic instabilities) but in some cases, a kinetic description is required (microscopic, or kinetic instabilities). The linearization of the equations gives the dispersion relations, i.e., the frequency and growth rate of the perturbations as a function of wavenumber, and are extremely useful to understand the mechanisms responsible for the development of the instabilities. However, this approach can describe only the linear stage of the development of the instability and cannot predict how the instability saturates. Numerical simulations based on fluid or kinetic (e.g., Particle-In-Cell simulations) transport equations must be used to study the evolution from the linear stage to the saturated stage of the instability but the interpretation of numerical results is not always straightforward. The best approach to fully understand the physics of the instabilities is, therefore, to combine the derivation of the dispersion relations by linearization of the equations, with numerical simulations.
plasma devices are complex physical systems, and it should be kept in mind that before trying to develop fully predictive engineering models of these devices, it is important to build simplified or reduced models at the right level of description and approximation. As stated in Ref. 92, “Every good model starts from a question. The modeler should always choose the correct level of detail to answer the question.” This is clearly true for the derivation of analytical dispersion relations but also applies to numerical simulations.
This section is organized as follows. The principles of the derivation of fluid dispersion relations in the context of plasmas are described in Subsection III A. The derivation of kinetic dispersion relations is only briefly discussed in Subsection III B. The kinetic dispersion relation of an plasma in the conditions of a Hall thruster, leading to the electron cyclotron drift instability will be illustrated in Sec. V. In Subsection III C, we recall the basic principles of Particle-In-Cell Monte Carlo Collisions (PIC-MCC) simulations that are often used to study plasmas and their instabilities. Finally, Subsection III D shows some examples of instabilities predicted by PIC-MCC simulations and that can be understood on the basis of fluid dispersion relations.
A. Theoretical description of instabilities—Fluid dispersion relations
In this section, we describe some aspects of the derivation of the fluid dispersion relations of a low-temperature partially magnetized plasmas based on Refs. 2 and 93–95. We use the notations of Fig. 1, where is in the direction, in the direction, and in the direction. The plasma density gradient is parallel to the electric field, in the direction.
We look for solutions of Eqs. (11)–(14) above, of the form , where is the complex wave frequency and the wavenumber in the direction (i.e., ). , the real part of , is the wave angular frequency and , the imaginary part, is the growth rate. The perturbation grows when . In general, the equilibrium and perturbed density, and are function of the axial coordinate, and the full nonlocal theory should be used. For the long-wavelength perturbations, the local theory can be used, assuming the density gradients constant.
For , the solution of Eq. (21) is purely real. This corresponds to the stable, anti-drift mode.2,36,96 The term in the denominator, corresponding to the flow, is a possible destabilizing term, since, for example, when and , becomes negative and Eq. (21) has a positive imaginary root ( ) and the instability grows. This instability, which occurs due to the destabilization of the anti-drift mode by the electron flow, is called the gradient-drift instability or the collisionless Simon–Hoh instability or the modified Simon–Hoh instability.35,36 In the collisionless case, the instability is reactive, i.e., the dispersion relation is real, and the complex roots come in complex conjugate pairs. This is an example of negative energy mode in flowing plasmas and fluids when the instability is caused by the coupling of positive and negative energy modes (a plasma wave is said to have negative energy if energy is removed from the plasma when the wave is excited; energy is added to the plasma in the case of a positive energy mode). Negative energy modes may become unstable solely due to dissipation. As discussed by Sakawa and Joshi,36 the dissipative Simon–Hoh instability was studied by Simon,97 Hoh,98 and Thomassen99 in a weakly ionized, inhomogeneous, collisional, magnetized plasma column and was found to be unstable when the density gradient and electric field were in the same direction (i.e., , as in the collisionless case). An important point in the Simon–Hoh instability is that the electron drift is much larger than the ion drift. This difference in electron and ion drift can be due to the weak magnetization of ions or to collisions.
Typically, for unmagnetized ions the ion azimuthal drift is neglected. In the case of weak magnetization , the ion drift can be estimated as100 where is the modified Bessel function (see Sakawa et al.100 for the derivation of this expression, which is valid only in a slab geometry and does not include centrifugal effects).
These expressions are valid when the plasma current is neglected, i.e., it does not modify the magnetic field and . These expressions also show that for a uniform magnetic field, electron temperature gradients and generally electron temperature dynamics are not coupled to the electron density directly (unless the electron temperature changes the density via ionization). For a nonuniform magnetic field, the compressibility of the electron diamagnetic flow in Eq. (29) may involve temperature evolution directly. In that case, the dispersion relation (25) becomes more complex and involve the equilibrium temperature gradients, as described in Frias et al.93 Among other effects, the temperature gradients affect the condition for the instability (27) and may result in the appearance of unstable modes in the direction opposite to the drift.93
Because of the condition (29), in the case of a uniform magnetic field, the diamagnetic drift does not result in the frequency Doppler shift in the electron dynamics, as it occurs due to the , and does not result in additional terms in the Simon–Hoh dispersion relation (24), contrary to the statement in Ref. 102. We note also that the condition , obtained in Ref. 102, is equivalent to the condition , which is not possible in the fluid theory but requires a full kinetic treatment.
The derivation of the dispersion relations (21) and (25) was made neglecting the inertia terms in the electron momentum equation. Also, the approach based on the standard momentum equation assuming scalar pressure cannot describe situations where the wavelength of the instability becomes close to the electron Larmor radius (i.e., when and a kinetic approach is necessary.
For very large values of , the general form of Eq. (30) gives the ion sound mode or for finite Debye lengths.
Note that Eq. (30) can be written in dimensionless units where the frequencies are normalized by , the lengths are normalized by , and the velocities are normalized by . If we write , , the roots of Eq. (30) depend on three parameters: , , and . If finite Debye length is taken into account, another parameter is (or, equivalently, ).
We show in Figs. 9 and 10 some examples of solutions of the dispersion relation (30) for different values of and , with and and a uniform magnetic field. In Fig. 9, the dimensionless diamagnetic drift velocity is fixed and the drift velocity is varied. In Fig. 10, is fixed and is varied. The corresponding numerical values of , and other parameters are given in the figure captions, for a particular case corresponding to argon, a magnetic field intensity of 50 mT, and an electron temperature = 5 eV.
We see in Fig. 9 that the maximum growth rate is obtained for decreasing values of the wave number, i.e., increasing wavelength, when increases at constant . It is equivalent to say that the wavenumber at maximum growth rate increases with the electric field for a constant plasma density gradient defined by . In Fig. 10, the wavenumber at maximum growth rate decreases with decreasing diamagnetic drift at constant drift. Long wavelengths are, therefore, obtained for low values of the diamagnetic drift or large values of the drift. At the limit of zero density gradient or zero diamagnetic drift ( or , the lower-hybrid mode is destabilized by drift and collisions only, in coherence with the simplified relation (34).
From the general dispersion relation (30), one can calculate, for each value of and , the maximum growth rate , the wave angular frequency at the maximum growth rate, , and the wavenumber at the maximum growth rate, or the wavelength at maximum growth rate. The corresponding dimensionless values, , , and are displayed in the form of contour plots as a function of and , in Fig. 11 (similar plots were presented in Ref. 105). On each of these plots, the Simon–Hoh or gradient drift instability is obtained in the lower right part of the figure (long wavelengths, lower angular frequencies) and the lower hybrid mode in the upper left part (shorter wavelengths, higher frequencies). The resistive mode is obtained for (not seen on the plots because of the log scale), while the lower-hybrid mode destabilized by density gradient and collisions is obtained for .
The instabilities discussed in this section are closely related to Farley–Buneman and gradient-drift instabilities in ionospheric plasmas.106
The dispersion relations above have been obtained assuming perturbations in the ( ) direction only. More comprehensive treatment of the fluid dispersion relations in low-temperature partially magnetized plasmas can be found in the literature, which consider effects such as electron or ion flow in the direction,93 electron temperature perturbations,93 sheath effects,107 etc.
Finally, as mentioned in the introduction of this section; dispersion relations give information on the development of the instabilities in the linear stage only. Analyzing the growth of the instability and the saturation mechanisms is more challenging and requires the use of full fluid or kinetic models. Examples of fluid description of the evolution of the instabilities based on the BOUT++ code108 can be found in Refs. 2 and 109.
B. Kinetic dispersion relations
The derivation of the fluid dispersion relations above shows the important role of plasma density gradient and magnetic field gradient in the development of instabilities in plasmas. Simulations show that instabilities can also develop in the absence of plasma density and magnetic field gradients in low-temperature partially magnetized plasmas. These are very large wavenumber modes corresponding to wavelengths close to the electron gyroradius. In these conditions, the description of the instability must take into account the electron cyclotron orbits and requires a kinetic treatment. The dispersion relation in that case is obtained from the linearization of the Vlasov equation for electrons and of the continuity and momentum equations of the unmagnetized ions (same as in Subsection III A).
The full derivation of the kinetic dispersion relation of low-temperature partially magnetized plasmas is beyond the scope of this tutorial. We refer to Refs. 110–116 and references therein for this derivation. The expression of this kinetic dispersion relation is given and discussed in Sec. V on instabilities in Hall thrusters.
C. Particle-In-Cell Monte Carlo Collisions simulations
In Particle-In-Cell Monte Carlo Collisions (PIC-MCC) simulations, the trajectories of a large representative number of particles (superparticles) are followed in phase space under the influence of the electric forces due to the self-consistent electric field and external magnetic field, and of collisions. The electric field is calculated at each time step from the charged particle densities on this grid, deduced from the positions of the superparticles. The electric forces are then calculated on the grid and the equations of motion of the superparticles are integrated over time. A Monte Carlo module is used to decide which particles are undergoing collisions during the time step, the nature of the collisions, the change in velocity, etc. We will not discuss here in detail the principles of PIC-MCC and simply refer to standard text books117,118 and review papers.119–122 Three-dimensional particle models are still very computationally intensive, especially if one wants to resolve the large plasma frequencies, small Debye lengths, and long time to reach steady-state of Hall thrusters or magnetron discharges. Most of the particle simulations of the plasmas of Hall thrusters and magnetron discharges have been performed with 1D3V or 2D3V models, i.e., one or two dimensions in space, three dimensions in velocity, and we will limit our discussion of this article to these models (we note also the recent attempts at reducing the computational cost of 2D PIC simulations of Hall thrusters by introducing the concept of pseudo-2D PIC scheme123,124). As said above, the dominant instabilities develop in the azimuthal direction because of the large difference between electron and ion drift velocities in this direction. Therefore, the 1D and 2D models must include the azimuthal direction. The 1D models are purely azimuthal, and the 2D models can be axial–azimuthal or radial–azimuthal. This will be illustrated in Secs. V and VI.
D. Example of evidence for rotating instabilities in E × B plasma devices
The aim of this subsection is to show a few examples of evidence for the presence of rotating azimuthal instabilities in low-temperature partially magnetized plasmas, drawn from simulations, and which illustrate the predictions of the fluid dispersion derived above. In particular, we will see examples where PIC MCC simulations predict the formation of long wavelength Simon–Hoh instabilities due to density gradient and electric field being in the same direction, and small-scale lower hybrid instabilities driven by density gradient and electron–neutral collisions. The general context are the experiments of Raitses et al.125,126 on a Penning discharge, but the conditions of the simulations are not necessarily close to those of the experiments. A discussion of instabilities under conditions closer to a real device will be given in Secs. IV–VI, with the example of Hall thrusters and magnetron discharges.
We start with the Penning discharge schematically represented in Fig. 12. The principles and conditions of operation of this discharge are indicated in the caption to Fig. 12. In these conditions, rotating plasma non-uniformities (“spokes”) in a plane perpendicular to the magnetic field are observed in the experiments with both fast imaging techniques (CCD camera) and Langmuir probes measurements. Rotating spokes have been observed in many other plasma devices including standard Hall thrusters,129–133 cylindrical Hall thrusters,134 wall-less Hall thrusters,135 anode layer Hall thrusters,136,137 low power dc or rf as well as high power pulsed magnetron discharges,138–145 and, of course in magnetized plasma columns dedicated to the study of turbulent phenomena.37–39,146,147
A number of articles have also been devoted to modeling and simulation of rotating spokes in recent years, with kinetic approaches (Particle-In-Cell Monte Carlo Collisions, PIC-MCC)26,105,127,128,148–154 or fluid155 and hybrid fluid–particle simulations.156–158 An example of results of particle simulations of the Penning discharge of Fig. 12 is shown in Fig. 13 while Fig. 15 displays some high-speed camera images of a rotating spoke in a wall-less Hall thruster.135
Figure 13 shows an example of PIC-MCC simulation results of the Penning discharge of Fig. 12 taken from the article of Powis et al.128 The conditions of the simulations are indicated in the caption to this figure. The figure illustrates the formation of a spoke in argon at 1 mTorr, and its rotation in the direction at a velocity at mid-radius of 4.9 km/s. Simulations have also been performed in a purely collisionless situation and compared with this collisional case with self-consistent ionization (in the collisionless case, positive ions are injected in the central region of the discharge).
The main results of the paper by Powis et al. can be summarized as follows. The PIC-MCC simulation predicts the formation and rotation of a spoke under both the collisionless and collisional conditions, with similar rotation velocities. Anomalous electron transport to the anode can be analyzed by looking at the time-averaged profiles of the electric field and plasma density. Figure 14 shows the time-averaged radial distribution of the electric potential, radial electric field, and electron density in the collisionless and collisional case. We see that the injection of electrons and the trapping of electrons by the magnetic field tend to lower the potential in the plasma center.
Equation (2) was derived for situations where electron transport across the magnetic field was due to collisions with neutrals. The index “eff” in Eq. (35) indicates that we consider an effective electron collision frequency to represent the effect of instabilities on cross field transport in the same way as in the purely collisional case.
The effective mobility and Hall parameter and are obtained simply by equating the expression (35) of the mean cross field velocity [with and from Fig. 14] to the mean velocity deduced from the PIC-MCC simulation (a similar approach was used in Ref. 105) The radially (and time) averaged effective collision frequency calculated by Powis et al. in the conditions of Figs. 13 and 14 was on the order of 108 s−1 both in the collisionless and collisional cases. For comparison, the real electron–neutral collision frequency in the collisional case (1 mTorr pressure) was about ten times smaller, indicating that electron cross field transport is also dominated by fluctuations at 1 mTorr. The calculated effective Hall parameter was between 13 and 16 in both cases.
One of the conclusions of the article by Powis et al. is that, since similar rotating spokes are obtained in the Penning discharge model for both collisional and collisionless cases, ionization is not a necessary condition for the existence of spokes. We will see, however, in Sec. IV that under magnetron discharge conditions, ionization may play a major role in the physics of a rotating spoke and that an instability triggered by the Simon–Hoh criterion can evolve into an ionization wave. 3D PIC-MCC simulations149 performed in the conditions of the wall-less Hall thruster of Fig. 15 also suggest that the rotating spoke can be associated with an ionization wave coupled with a periodic depletion of the neutral atom density (breathing mode, see Subsection III A).
Tyushev et al.153 studied a Penning discharge under conditions similar to those of Powis et al., with collisional electrons and collisionless ions. By varying the power input by the electron beam into the discharge, they could control the radial profile of the time-averaged electric potential.
In the case of a deep potential well, they found a spoke regime as in the results of Powis et al.,128 Fig. 13. When the potential profile was flatter, they observed a transition into a rotating small scale spiral arm with m , as illustrated in Fig. 16(a). This is a clear illustration of a transition from a long-wavelength Simon–Hoh instability (spoke), where the instability is triggered by the density gradient and radial electric field being in the same direction, to a small-scale lower hybrid instability driven by density gradient and electron–neutral collisions. Similar conclusions were drawn in Ref. 105, where the 2D PIC-MCC simulation was performed in an axial–azimuthal geometry rather than in a radial–azimuthal geometry. In these simulations, the profile of the electric potential was controlled by the magnetic field intensity.
An example from this reference is shown in Fig. 16(b) in the conditions of small-scale instabilities. Lucken et al.159 also found the formation of rotating spiral structures with in 2D PIC-MCC simulations under conditions where the plasma was generated by radio frequency electron heating in the direction parallel to the magnetic field, i.e., perpendicular to the simulation domain. This is illustrated in Fig. 16(c).
In all three models,105,153,159 the effective Hall parameter was found to decrease, i.e., the anomalous transport was increasing, from the spiral regime (low magnetic field or quasi-flat electric potential) to the spoke regime (larger magnetic field or deeper potential well), with a saturation at large magnetic field in Refs. 105 and 159.
IV. PHYSICS OF HALL THRUSTERS AND MAGNETRON DISCHARGES
In this section, we describe in more details the physics and instabilities of two important plasma devices, the Hall thruster and the magnetron discharge.
A. Hall thrusters
The plasma of a Hall thruster (Fig. 17) is generated in a channel between two concentric dielectric cylinders. The anode is at one end of the channel. Xenon (or krypton) is injected through the anode. An electron emissive cathode is placed outside the channel exhaust. A radial magnetic field, with a maximum (about 20 mT) close to the exhaust plane, is generated by coils or magnets and a magnetic circuit. A voltage of 300 V is applied between cathode and anode under nominal operations. The xenon mass flow rate, and radii of the inner and outer dielectric cylinders are about 5 mg/s and 3.5 and 5 cm for a 1 kW thruster. Because of the lowering of the axial electron mobility in the region of maximum magnetic field, the axial electric field in that region is expected to increase.
The purpose of a Hall thruster is to extract ions from a plasma and generate thrust without using biased grids like in usual ion sources, but by creating a large accelerating field within the plasma. The Hall thruster is a very interesting and unique plasma device where a potential drop of several hundred volts can indeed be generated in a quasineutral plasma. The electric field in this potential drop can be as large as 400 V/cm over a distance of about 1 cm, much larger than the Debye length. This is quite remarkable since one could expect the cross field electron mobility to be greatly increased by instabilities and turbulence, leading to a collapse of the electric field.
One thing that helps maintaining a low electron conductivity in the acceleration region and outside the channel is the fact that the flow of neutral atoms coming from the anode side is practically fully ionized by the energetic electrons in the ionization region upstream of the channel exhaust [see the location of the ionization region in Fig. 17(b)].
Although instabilities certainly play an important role in the enhancement of cross field electron transport in a Hall thruster, they are not the only cause of the larger than expected electron mobility. Electrons interactions with the channel walls also contribute to the increase in electron effective mobility inside the channel. The scattering of fast electrons by the channel walls and the secondary electron emission resulting from this interaction are also important causes of cross field electron transport. The interactions of electrons with the channel walls also have important impact on the electron energy balance in the channel. There is a considerable literature on the effect of electron–wall interactions on the conductivity in Hall thrusters and a discussion of these effects is beyond the scope of this paper. Discussions and references on this question can be found in Kaganovich et al..1
Controlling the electron conductivity in a Hall thruster is an important issue because the efficiency of the thruster decreases if the electron current entering the channel increases5,7 (i.e., if the electron conductivity increases in the exhaust region). The operations and performance of a Hall thruster are extremely sensitive to the distribution and axial profile of the radial magnetic field which is still empirically optimized (models with predictive capabilities are not available). This sensitivity is largely due to the fact that the magnetic field intensity profile controls the instabilities and cross field electron transport. For example, the nature of azimuthal instabilities may be different in different parts of the thruster, as illustrated in Fig. 19, which shows schematic axial profiles of the radial magnetic field, axial electric field, and plasma density. The plasma density decreases outside the channel and toward the anode. On the anode side of the thruster, the plasma density and axial electric field are in the same direction. This suggests that long wavelengths instabilities of the Simon-Hoh or gradient-drift type may be present in this region. Rotating spokes have indeed been observed in this region,130–134,161,162 and their presence depends on the magnetic field intensity. On the cathode side of the maximum magnetic field, the electric field and plasma density gradient are in opposite direction so a gradient-drift instability cannot develop in that region. We will see in Sec. V that a short wavelength kinetic instability, the electron cyclotron drift instability, and a fluid instability, the modified two-stream instability, can be present in the acceleration region of a Hall thruster.
Figure 20 shows the different types of instabilities that can develop in a Hall thruster. Rotating spokes, in the frequency range 5–20 kHz, electron cyclotron drift instabilities around 2–5 MHz and modified two-stream instabilities at 5–10 MHz form in the azimuthal direction and are directly related to the configuration, but axial instabilities such as the breathing mode and the ion transit time can also be present and sometimes coupled with the azimuthal fluctuations. An early review of Hall thruster oscillations can be found in Ref. 163. Although progress in understanding these oscillations has been made over the past twenty years and is described in Ref. 1, a fully self-consistent and predictive description of Hall thruster instabilities has yet to be achieved. We briefly summarize these instabilities below, and describe some of them in more details in Sec. V.
The breathing mode is an ionization instability related to a periodic depletion of the neutral density due to strong ionization of the neutral flow. This is one of the most common and most studied instabilities in Hall thrusters.6 This ionization instability can lead to large amplitude oscillations of the discharge current. The breathing mode is strongly related to the cross field electron transport and to the electron energy balance. Strong ionization by electrons trapped in the large magnetic field region and drifting in the direction induces a depletion of the neutral density in the ionization region. The neutral density depletion changes the momentum and energy balance of the electrons, leading to a decrease in electron conductivity and in the current of electron entering the channel. The ionization decreases until the flow of neutral atoms from the anode replenishes the neutral depletion. The oscillation frequency is therefore related to the time necessary for the neutrals to refill the ionization region. This is a low-frequency mode typically in the 5–20 kHz range. This qualitative description actually hides a very complex and non-linear physics that is still under discussion. Excitation of the breathing mode is affected by several interrelated mechanisms that depend in a complex way on the magnetic field configuration and are not easily quantifiable. These mechanisms include electron momentum and energy losses to the walls, anomalous electron transport, ion backflow and recombination on the walls and anode, etc. Attempts to untangle these complex physical mechanisms have been made in a number of papers based on 1D164–173 numerical models. A simple 0D predator–prey model has been proposed165,169,171,174 to describe the ionization–depletion–replenishment process but fails to identify the conditions for the instability. We will not discuss these research efforts further here and we refer to some early or more recent review articles1,175 and to recently published papers on this subject.172,173,176
Under certain conditions, the low-frequency breathing mode coexists with high-frequency oscillations in the range of 100–500 kHz. These high frequency oscillations correspond to the characteristic time-of-flight of ions in the channel and have been evidenced in hybrid Hall thruster models and in experiments.173,177–184 Using 1D fluid and hybrid models, Chapurin et al.173 identified two distinct regimes of breathing oscillations depending on the assumption on the electron energy losses to the walls. The regime with higher energy losses exhibits a low-frequency mode of about 14 kHz that coexists with high-frequency ion transit time oscillations around 150 kHz. In the second regime with low electron energy losses, pure breathing oscillations are observed.
B. Magnetron discharges
We have seen in Subsection II C that in a magnetron discharge, the magnetic field is generated by magnets placed behind the cathode (Fig. 21). This creates a point-cusped magnetic field. Electrons are mirroring at the cusp regions and are also electrostatically confined by the cathode sheath. The plasma forms in the region above the cathode, where the magnetic field is parallel to the cathode. This is an configuration with axial electric field and radial magnetic field. The magnetic field intensity decreases from cathode to anode (as in the channel of a Hall thruster). Magnetron discharges can be operated in dc or rf regimes, but also in a pulsed regime with very high-power densities (HiPIMS) at gas pressures below a few Pa. They are used in sputter deposition applications.
The ions are accelerated at high energies toward the target cathode in the quasi-collisionless sheath and sputter atoms from the cathode surface. These atoms then deposit on a substrate, forming a thin layer.
A number of papers report the presence of instabilities in magnetron discharges, which take the form of large-scale plasma non-uniformities rotating in the azimuthal direction and associated with rotating regions of enhanced light emission and ionization. These rotating non-uniformities were first detected in Hall thrusters by Janes and Lowder185 using Langmuir probes and were called spokes. More recent observations with high speed camera imaging in Hall thrusters130,133–135,186,187 as well as in magnetron discharges138,141,142,188–192 have shown that these rotating non-uniformities are associated with a strong light emission.
Interestingly, it has been shown in several articles142,190,192–194 that the spoke can rotate in the direction in low power dc magnetrons and in the direction in high power pulsed magnetrons used in plasma processing applications (HiPIMS) and in some Hall thrusters, at velocities more than ten times smaller than the drift velocity. The images of rotating spokes in a dc magnetron discharge obtained by Panjan and Anders194 and displayed in Fig. 22 show that the number of spokes increases with gas pressure. Stable plasma patterns with four and up to seven spokes were observed by these authors under rf (radio frequency) voltages with self-bias. At high enough pressure, collisional electron transport becomes dominant and the spokes disappear.195
An important issue that has been widely discussed in the literature on magnetron discharges15,139,192,196,197 relates to the electron heating mechanisms leading to the formation of these luminous regions, one conclusion being that electrons are not heated in the cathode sheath as in non-magnetized glow discharges but rather in the presheath, by Ohmic heating.
We will show in Sec. VI that particle simulations provide a new insight in the electron heating mechanisms in the spokes of magnetron discharges.
Space- and time-resolved Langmuir probe measurements of the plasma and floating potentials in dc magnetron discharges by Panjan and Anders194 have revealed the existence of strong electric fields parallel and perpendicular to the cathode surface, forming a double layer structure at the leading edge of the ionization zone. Figure 23 shows an example of measured plasma potential distribution in a plane-parallel to the cathode and 7.5 mm above it. A sharp drop of the potential of several tens of volts can be seen at the spoke front. The magnetron discharges in Figs. 22 and 23 operate in a low power dc regime, and the spokes rotate in the direction.
These measurements suggest that the double layer at the leading edge of the spoke plays a crucial role in the energization of electrons in the spoke. Held et al.145 measured the plasma potential, electron density, electron temperature, and electron energy probability function (EEPF) with Langmuir probes, as a function of time at a given location above the cathode, on the spoke path.
The sharp increase in the plasma potential when the spoke front reaches the probe, seen in the measurements of Held et al., Fig. 24) is consistent with the 2D potential distribution of Fig. 23 and is associated with a very fast increase and decay of the electron temperature. The electron density presents large amplitude oscillations as a function of time in the spoke region. Held et al.145 also measured the electron energy distribution at different times and observed a fast increase in the tail of this energy distribution at the spoke front, followed by a quick decay. We will see in Sec. V that the particle simulations of magnetron plasma are qualitatively consistent with these measurements.
V. MICROSCOPIC (KINETIC) INSTABILITIES IN HALL THRUSTERS
Hall thrusters and other plasma devices are complex because of the different non-linear and coupled physical mechanisms involved (instabilities, ionization, neutral depletion, and other interaction of charged particles with the neutral gas). As discussed in the introduction of Sec. III, rather than trying to deal with the full complexity of the problem it is often useful to simplify or reduce the problem and to decouple some of these mechanisms. This is illustrated in this section with the example of a Hall thruster, where we focus on the electron cyclotron drift instability, a kinetic instability that has been shown to develop in this device.
The existence of the electron cyclotron drift instability (ECDI) in the particle simulations of Hall thrusters has been demonstrated for the first time by Adam et al.198 in 2004 with a 2D PIC-MCC axial–azimuthal model including a fluid description of the neutral atom transport, ionization, and neutral atom depletion. The model was able to describe the azimuthal ECDI and the breathing mode mentioned above. The ECDI was characterized, in these simulations, by the presence of large amplitude, small wavelengths azimuthal oscillations of the azimuthal electric field, and plasma density in the acceleration region of a Hall thruster. The EDCI was then studied in more details by Ducrocq et al.115,116 with dispersion relations and a simpler 1D PIC-MCC simulation. An analysis and a parameter study of the 3D dispersion relation of the ECDI in Hall thrusters can be found in the paper by Cavalier et al..113 The dispersion relation for the ECDI had been derived and studied earlier,110,199–201 in the 1970s, in other contexts such as collisionless shocks in the magnetosphere.
In this section, we consider PIC simulations based on reduced models where the plasma is collisionless. This is a good approximation downstream of the ionization region. The PIC simulations contain the azimuthal direction and, to represent closed-drift, the boundary conditions are periodic in the azimuthal direction. In Subsection V A, the ECDI is described by a 1D PIC simulation in the azimuthal direction. The theory of the ECDI is outlined in Sec. V B. Additional details on the properties of the ECDI in Hall thrusters are discussed in Secs. V C and V D on the basis of 2D axial–azimuthal and radial–azimuthal PIC simulations, respectively. We conclude this section by some remarks on experimental evidence of the presence of the ECDI in Hall thrusters in Subsection V E.
A. 1D Particle-In-Cell simulations of instabilities in a Hall thruster
In 1D models, the particle simulation is performed in the azimuthal, direction (same notations as Fig. 1), i.e., electrons and ion trajectories are followed in the direction and the electric field is re-calculated at each time step in this direction, with periodic boundary conditions. The axial electric field , in the direction perpendicular to the azimuthal direction of the simulation and to the magnetic field is given and is supposed to be constant. represents the electric field of the acceleration region of the Hall thruster. The magnetic field intesity in the direction is also given as constant and represents the radial magnetic field of the Hall thruster. The simulation is collisionless and starts with given Maxwellian velocity distributions of electrons and ions. As the instability forms, anomalous transport across the magnetic field occurs and electrons can gain energy from the axial electric field. Since the electric field is fixed in the direction perpendicular to the direction of the simulation, the electrons constantly gain energy in this direction. In order to limit the electron mean energy, a finite dimension can be assumed in the direction of the electric field. The electron positions in the direction of the applied electric field can be tracked, and each electron that has traveled a given length along the electric field (the acceleration region) is removed from the simulation and replaced by an electron with the initial velocity distribution.
Figure 25 shows an example of result from a 1D Particle-In-Cell simulation, from Ref. 112. A short wavelength instability develops in the simulated azimuthal direction. An analysis of the wavelength of the instability shows that the azimuthal wavenumber closely satisfies the relation (cyclotron resonance) where is the azimuthal drift velocity and the cyclotron angular frequency. We can say that the instability is dominated by a cyclotron resonance and appears as a quasi-coherent mode. Higher modes with larger growth rates, defined by and > 1, can also be observed during the early stage of the instability and have larger growth rates (see Fig. 27) but the increase in electron temperature and nonlinear inverse cascade make the mode dominant in the nonlinear stage. We will see in Subsection V B that the instability observed in the simulation is consistent with the theory defining the electron cyclotron drift instability (ECDI).