Z Pinch Kinetics II - A Continuum Perspective: Betatron Heating and Self-Generation of Sheared Flows
Abstract
Adiabatic compression of a self-magnetizing current filament (a Z pinch) is analyzed via the adiabatic invariants of its constituent cyclotron and betatron motions. Chew-Goldberger-Low (CGL) models are recovered for both trajectories but with distinct anisotropy axes, about the magnetic field for cyclotron fluid and about the electric current for betatron fluid. In particular, betatron heating produces agyrotropic anisotropy which balances with gyrophase mixing. A hybrid CGL model is proposed based on the local densities of cyclotron and betatron orbits, then validated by numerical experiments. The relation between anisotropy and shear is explored by constructing the kinetic equilibrium of a flow expanded in the flux function. Flow as a linear flux function is simply bi-Maxwellian, while higher powers display higher-moment deviations. Next, weakly collisional gyroviscosity (magnetized pressure-strain) is considered in a forward process (forced flow) and an inverse process (forced anisotropy). The forward process phase-mixes flow into a simple flux function, freezing flow into flux and inducing anisotropy. In the inverse process, betatron heating-induced anisotropy self-generates a sheared flow to resist changes in flux. This flow, arising from momentum diffusion, is concentrated in the betatron region.
I Introduction
The Z pinch is a quasineutral cylindrical filament of electric current conducted through a plasma, concentrating equilibrium plasma pressure energy in proportion to the self-magnetic energy of the current. The Z-pinch equilibrium was originally identified in 1934 by Willard H. Bennett, with the assistance of Llewellyn Thomas, as a hypothesis to explain the high-voltage breakdown of gases Bennett (1934), and termed “the magnetic pinch effect” in 1937 by Lewi Tonks Tonks (1937). The pinch as a fusion energy concept was extensively studied in the early, classified stage of fusion research Artsimovich (1962); Bostick (1977). Many works on pinch theory appeared at the time of fusion research declassification in 1958, at which time Lawson Lawson (1958) and Budker Budker (1956) independently explained how the concept of Bennett’s pinch also describes magnetically self-focused charged particle beams. A critical parameter of magnetic self-focusing was identified, now called the Budker parameter, given by Linhart (1959)
(1) |
The manifold nature of Eq. 1 is emphasized, as various authors use the same parameter under different names. Here represents particles of species per unit length, the classical particle radius, species thermal velocity, species partial drift velocity, species partial current, species Alfvén limiting current Alfvén (1939), Larmor radius, skin depth, the pinch radius, species cyclotron frequency, species thermal crossing time, the Hartmann number of the flow pinch, and the factors are and .
Additional key parameters of charge neutralization fraction and relativistic -factor were noted in 1959 by Lawson Lawson (1959). The relations of Eq. 1 involving species velocity hold for both electrons and ions in the Lorentz frame where the two are counter-streaming and thus perfectly charge neutralized (for ). The relations of Eq. 1 are derived with commentary in the prequel Crews, Meier, and Shumlak (2024).
The Budker parameter measures the ratios of: temperature to drift-kinetic energy, plasma current to Alfvén limiting current, characteristic Larmor radius/skin depth to pinch radius, the species dynamical Hall parameter, and the magnitude of electromagnetic to viscoresistive hydrodynamic forces. Perhaps most importantly, measures the magnetization of orbits. When , a non-negligible part of the canonical ensemble form cyclotron orbits, initially in the periphery where magnetic flux density is high. The remaining particles in the current-dense core follow betatron orbits, a radially bouncing, axis-encircling drift motion associated with magnetically self-focused but “unmagnetized” streams. The prequel Crews, Meier, and Shumlak (2024) shows that the ensemble-averaged canonical momentum passes from an unbound state to a bound state through , analogous to the positive and negative energies of gravitational or electrostatic binding Arnold (2013); Crews, Meier, and Shumlak (2024).
The limit is the negligible self-field neutralized beam limit (typically with ), while is the magnetohydrodynamic (MHD) limit Finkelstein and Sturrock (1959). For neutralized beams, MHD modes are thought to be stabilized by betatron orbit effects Bennett (1955); Finkelstein and Sturrock (1959), with a residual stabilizing effect when in the “large ion Larmor radius” (LLR) regime Haines (2001) (). Interestingly, the terms “LLR Z pinch” and “neutralized ion beam above the Alfvén limit” are synonymous in the perfectly neutral Lorentz frame but not synonymous in the laboratory frame, distinguished by the laboratory-frame axial velocity of the ion betatron orbits. The term “neutralized beam above the Alfvén limit” is usually applied to electrons in the relativistic regime Santos et al. (2018). Betatron resonance is receiving increasing attention in the context of plasmas forced by lasers near the critical density Arefiev et al. (2024).
The frame-dependent syzygy of neutralized ion beams and LLR Z pinches in the regime is closely related to the notorious Z-pinch beam-target fusion mechanism Vikhrev and Korolev (2007). Experimental Thomson scattering measurements on the Fusion Z-Pinch Experiment (FuZE) Goyon et al. (2024) suggest that the deuterium Budker parameter at the time of neutron radiation, while for the electrons . Neutron isotropy measurements on FuZE have been found to be inconsistent with typical Z-pinch beam-target fusion mechanisms Mitrani et al. (2021), suggesting LLR Z-pinch laboratory-frame ion dynamics and thereby motivating the present study of kinetic effects in flow pinches.
In the prequel, the distribution of orbits was studied in the transitional magnetization regime (i.e., moderate ion Budker parameters, or the LLR regime). It was found that plasma species density is decomposable into betatron and cyclotron orbital densities, . The net density is fully composed of betatron orbits as , fully of cyclotron orbits as , and partially of both at intermediate radii within a transition layer through which current conduction switches from betatron to diamagnetic cyclotron conduction. By considering the canonical distribution (the Bennett pinch), the location and thickness of this transition layer relative to the pinch radius was found to depend only on , and lies close to the pinch radius in the LLR regime. This kinetic description of the Bennett pinch is a diffuse example of Kotelnikov’s theory of sharp diamagnetic layers in the Gas Dynamic Trap Kotelnikov (2020). Some current conduction by betatron orbits appears to be a necessary element of plasma configurations Timofeev, Kurshakov, and Berendeev (2024).
In this second work on Z-pinch kinetics, we apply the results of the prequel to the topics of collisionless adiabatic compression, kinetic equilibrium, and self-organization of sheared flows in the Z pinch. Collisionless momentum and heat fluxes in the intermediate magnetization regime are of particular interest to flow Z-pinch dynamics because Braginskii-type closures Braginskii (1965) do not describe the betatron-orbiting part of the plasma column even on moderately collisional timescales. On the weakly collisional timescales, phase-space mixing produces effective dissipation. While this allows hydrodynamic intuition to be applied somewhat to collisionless dynamics Du et al. (2020); Bandyopadhyay et al. (2023), kinetic effects can lead to unexpected phenomena such as flow-organizing and flow-generating viscosity Del Sarto, Pegoraro, and Califano (2016); Del Sarto and Pegoraro (2017).
Collisionless momentum fluxes are often studied in the context of the magnetized pressure-strain (or Pi-D) interaction Yang et al. (2017), coinciding in the collisional limit with gyroviscosity Kaufman (1960). The collisionless viscosity associated with magnetized shear flows is observed to be an important element of astrophysical plasma turbulence at the ion scale Hellinger, Petr et al. (2024), for example in the magnetosphere Bandyopadhyay et al. (2020) and solar wind Pezzi et al. (2019). A thorough discussion of the pressure-strain interaction can be found in a series of articles by Cassak and Barbhuiya et al Cassak and Barbhuiya (2022); Cassak, Barbhuiya, and Weldon (2022); Barbhuiya and Cassak (2022). Because LLR Z pinches are themselves on the ion inertial scale, flow pinch physics naturally parallels the astrophysical studies. Notably, the phase space signature of the unmagnetized pressure-strain interaction has recently been studied and found to be closely related to the field-particle correlation Conley et al. (2024), suggesting a fruitful line of investigation for similar relationships in the kinetic theory of magnetized flows.
This work makes several novel contributions to Z-pinch physics which should be of interest to the general fusion community, including:
- •
-
•
A class of kinetic equilibria where axial flow is a power series in the flux function, demonstrating increasingly non-thermal features with higher powers in the flux (Section IV),
-
•
A study of dynamic kinetic gyroviscosity on weakly collisional timescales, in which gyro-phase mixing rapidly brings forced shear flows towards kinetic equilibrium (Section V.2.2),
-
•
Demonstration of spontaneous shear flow generation due to flux compression of the betatron fluid, exhibiting reactive resistance to rapid changes in magnetic flux (Section V.3),
-
•
An equivalence between the characteristic anisotropy of an Alfvénic sheared flow with the Budker parameter, indicating increased kinetic-instability-mediated viscosity in the low ion Budker regime (Section V.5),
These results are developed as follows. First, Section II considers conditions for kinetic processes faster than the thermalization time, Section III develops a model for the anisotropy produced by fast compression, and Section IV presents a theory of kinetic flow-pinch equilibrium. Collisionless gyroviscosity (magnetized pressure-strain interactions) is discussed in Section V, while Section VI concludes with a numerical validation of the anisotropy model.
II Kinetic dynamics in the flow Z pinch
The compressibility, transport properties, and stability of plasmas essentially depend on the ion distribution function, which in turn is governed by kinetic processes. Of particular interest then is the ion relaxation rate , which is the rate at which ions approach a Maxwellian distribution from intraspecies scattering. The ion relaxation rate is proportional to their density and inversely proportional to temperature to the 3/2 power Braginskii (1965),
(2) |
with the plasma frequency, the Coulomb logarithm, and the ion plasma parameter. The Z pinch may be heated by adiabatic compression to increase the ion density and temperature Shumlak (2020). Importantly, the ion relaxation rate varies only logarithmically during compression along an adiabat, as, if is a reference relaxation rate, one can show that
(3) |
where is specific entropy, the specific heat ratio for monatomic ions, and is the specific heat at constant volume given the ion gas constant for Boltzmann’s constant.
Equation 3 shows that the relaxation rate is approximately constant in an adiabatic process. The adiabatic approximation is accurate in the collisional limit, but is modified by collisionless orbits and shock wave dynamics. Adiabatic processes include transient compression as induced by rising discharge current Ryutov, Derzon, and Matzen (2000) and steady flow compression as discussed by Morozov and Solov’ev Morozov and Solov’ev (1980). Adiabatic flow-pinch compression was recently discussed by Turchi and Shumlak Turchi and Shumlak (2023), and ideal flow-pinch thermodynamics was recently studied by the authors Crews et al. (2024).
Classical Z-pinch transport physics depends on the magnetization of ion orbits and the ion relaxation rate. Ion orbit magnetization is measured by the ion Budker parameter , while dependence on the ion relaxation rate is measured by dimensionless parameters including the ion Hall parameter with the characteristic cyclotron frequency, and the ratio of the plasma dynamical frequency to the relaxation rate, . The dynamical frequency is the inverse of the dynamical timescale of interest, such as the plasma flow-through time (i.e., with plasma length and axial fluid velocity), the flux compression timescale, or the growth rate of instability modes, which is typically estimated by a characteristic frequency (e.g., Alfvén transit rate for MHD modes or plasma frequencies for microinstabilities).
The magnetized transport theory of Braginskii Braginskii (1965), and extensions thereof Hunana et al. (2022), is valid in the regime , , and . Important corrections may be determined for the flow Z pinch. For instance, as a unity- plasma, orbit magnetization is not radially uniform. The experimental value of suggests that of the total radial ion inventory follow betatron trajectories in a near-axis transitional magnetization layer Crews, Meier, and Shumlak (2024), within which radius magnetized orbit theory is not correct. Modification to gyroviscosity is a particularly important correction for the flow Z pinch. Construction of a transport theory valid for radii in the transitional magnetization layer is an open problem. Meier and Shumlak have proposed empirical adjustments to avoid singularities in the Braginskii transport coefficients around the axis Meier and Shumlak (2021).
When dynamical timescales are faster than the collision time () Braginskii closures are not valid for the cyclotron orbits either because adiabatic dynamics lead to non-Maxwellian features beyond the closure model. In this situation, fluid elements no longer conserve specific entropy but instead conserve multiple invariants associated with the adiabatic invariants of their constituent particles. Multiple adiabatic invariants lead to anisotropic dynamics.
Experimentally relevant collisionless dynamics may arise during steady-flow compression at sufficiently high velocities or small length scales, transient flux compression events, and MHD instability growth at large ion Hall parameter. In addition, the consequences of these collisionless dynamics persist when the plasma lifetime is shorter than the relaxation time. The plasma lifetime may be considered the shorter of either the reactor flow-through time or the MHD instability growth time. Collisional processes can also interact in feedback with collisionless dynamics, such as the case of microinstability-enhanced resistivity Ryutov (2015).
Figure 1 shows the parameter space of the flow Z pinch in terms of ion temperature and density alongside calculated fusion gain contours Shumlak, Meier, and Levitt (2024). The filled contours show the fusion gain, while the colored lines show the ion relaxation time. The dashed lines show adiabatic trajectories, i.e. paths followed by the state of the plasma as it is heated by adiabatic compression. The black and green lines show fusion gain near and close to break-even level, and respectively.

Figure 1 shows that collisionless dynamics are relevant to breakeven fusion conditions in flow Z pinch plasmas, in this case where the ion relaxation time is s and thus comparable to the plasma lifetime and dynamical scales of interest (i.e., flow-through time, instability growth time, and compression time). Importantly, the ion relaxation time on an adiabat has the significance of the maximum time over which compression is singly adiabatic. Compression faster than the ion relaxation time, which we refer to as collisionless flux compression, produces anisotropic dynamics with a variety of interesting consequences.
III Adiabatic model of anisotropy in a compressing Z pinch
Dynamical processes on collisionless timescales generate pressure anisotropy, thus changing plasma equilibrium, momentum and heat transport, and kinetic stability Bott, Cowley, and Schekochihin (2024). The focus of this section is on pressure anisotropy generated by the dynamical process of collisionless flux compression. A model is developed which accounts for the transitional magnetization of pinch orbits from the current-dense center to the magnetized periphery, and applied in later sections to consider collisionless modifications to cross-field viscosity.
Compression is isotropic and collisional when slower than the relaxation time (see Braginskii Braginskii (1958), page 260, for a lucid description), and is adiabatic with respect to a single invariant, namely the specific entropy of the fluid elements. On the other hand, collisionless trajectories are associated with multiple adiabatic invariants, and collisionless compression which preserves these invariants can be described by a multiply adiabatic theory. Multiply adiabatic collisionless compression may be an inaccurate description in the presence of non-adiabatic collisionless processes such as shocks, radiation, and resistivity.
The set of single-particle adiabatic invariants depends on the particle orbit magnetizations. Well-magnetized particles on cyclotron trajectories conserve their magnetic moments and their angular momentum (or mirror invariant in a slab geometry), leading to doubly adiabatic compression described by the Chew-Goldberger-Low (CGL) equations Chew, Goldberger, and Low (1956). Magnetic moment conservation produces gyrotropic anisotropy, meaning the anisotropy symmetry axis is along magnetic field lines. In contrast, the adiabatic invariants of betatron trajectories lead to agyrotropic anisotropy under compression, meaning the anisotropy is symmetric about the electric current lines.
III.1 The CGL equations for fluid elements of cyclotron trajectories in a cylinder
Doubly adiabatic CGL theory describes collisionless compression of magnetized plasma, and predicts different degrees of parallel and perpendicular heating due to magnetic flux compression Chew, Goldberger, and Low (1956). The classic CGL model consists of two invariants, one for each pressure component
(4a) | |||||
(4b) |
in contrast to the singly adiabatic (conserved specific entropy) invariant .
Equations 4 hold in the cylindrical Z pinch for fluid elements which are made up primarily of cyclotron trajectories. This is demonstrated simply as follows. Suppose that a plasma fluid element is characterized by three adiabatic invariants, namely specific magnetic moment , specific angular momentum , and specific magnetic flux (from Alfvén’s law)
(5a) | |||||
(5b) | |||||
(5c) |
where is the parallel thermal velocity and the perpendicular one. Then substituting the parallel and perpendicular pressures ( and ) for the thermal velocities and combining Eqs. 5 yields exactly Eqs. 4, the CGL equations.
Equations 5 clarify the double-adiabatic CGL model in cylindrical geometry as consisting of the magnetic moment plus the angular momentum invariant, versus the Cartesian picture in which the second invariant is the mirror bounce action along a field-aligned path of length . The singly adiabatic model (invariant specific entropy) arises in the collisional limit . Observe that the CGL model is truly triply adiabatic considering the necessity of Alfvén’s law to obtain the CGL equations and the thermodynamic character of the specific magnetic flux Crews et al. (2024).
Compression in the Cartesian slab model and in the cylindrical pinch are distinguished by the Jacobian appearing in Alfvén’s law Turchi and Shumlak (2023). Namely, the Cartesian slab has while the cylinder has , a subtle difference which can lead to opposite expectations. That is, combining Eqs. 4 with these frozen flux invariants produces the following equations,
Slab, | (6a) | ||||
Cylinder, | (6b) |
Cartesian slab anisotropy is simply inversely proportional to the flux compression ratio, (with ) between an isotropic initial state and an anisotropic compressed state . In the same vein, anisotropy in the cylinder follows from Eq. 6b as
(7) |
where is the axial current enclosed up to radius .
Thus, the anisotropy produced by flux compression of cyclotron trajectories in a Z pinch is the product of two factors: the current amplification ratio and the ratio of radii (alternatively, the product of the flux and area compression ratios). Flux compression in the Cartesian slab always leads to the anisotropy orientation , but the Z pinch will easily display the orientation if current does not increase by a greater fraction than the radius. This azimuthal over-heating can happen because angular momentum conservation competes with magnetic moment conservation, and has important consequences for the resulting radial force balance.
III.2 Adiabatic invariants of betatron orbits and CGL model of the betatron fluid
To summarize the preceding, fluid elements composed of cyclotron trajectories display two adiabatic invariants (Eqs. 5), magnetic moment and angular momentum, and an additional frozen-in flux invariant. In the cylinder, magnetic moment is the invariant associated with radial and axial bounce of the cyclotron orbit, and the angular momentum with periodic azimuthal orbits.
Idealized betatron trajectories display invariants closely related to those of the ideal cyclotrons with a crucial difference: the lines of electric current are the principal symmetry axis rather than the magnetic field lines. Ergo, fluid elements composed of betatron trajectories should follow the same CGL equations as the cyclotrons, but the notions of parallel and perpendicular refer to the electrical current axis () instead of the magnetic field lines, which we demonstrate below.
The adiabatic invariants of an ideal betatron orbit consist of its radial bounce action , its axial bounce action , and its azimuthal angular momentum (or action of the periodic azimuthal motion). These invariants are inferred from the ideal betatron orbit (cf. the prequel Crews, Meier, and Shumlak (2024)), which we present here for self-completeness.
The radial potential energy of a charged particle in the vicinity of the Z-pinch axis is
(8) |
with the z-component of canonical momentum, provided that . Expanding the vector potential to around the Z-pinch axis yields a harmonic potential describing a radial bounce with spring constant . The axial bounce motion follows from conservation of canonical momentum. These radial and axial bounce motions are expressed as
(9a) | |||||
(9b) |
The axial momentum fluctuation amplitude is the potential momentum at the radial turning point and is the betatron frequency with . The radial and axial bounce motions contribute substantially to temperature in the species drift frame.
The radial bounce invariant follows as the area of the phase-plane orbit,
(10) |
or the product of the radial bounce amplitude and the radial velocity amplitude . The radial bounce action is similar in form to the angular momentum or mirror bounce invariants, i.e. velocity times length, as a simple harmonic motion.
The axial bounce invariant is found by transforming to the bounce-averaged () axial drift frame, i.e. to the oscillation-center coordinates . The oscillation-center trajectory is
(11a) | |||||
(11b) |
and the axial bounce invariant as the area of the oscillation-center phase-plane ellipse,
(12) |
Equations 10 and 12 hold approximately for non-ideal betatron orbits, in a similar way that magnetic moment invariance is taken to hold for cyclotron orbits, although it is truly conserved only up to guiding-center corrections. We also consider axis-encircling betatron orbits to be approximately described by Eqs. 10 and 12 with the addition of an angular momentum invariant, , because betatron orbits with azimuthal momentum may be approximated as an in-plane bounce orbit that rotates around the -axis Crews, Meier, and Shumlak (2024).
We now specialize to ion trajectories because their orbits are much larger than the electrons. The characteristic betatron frequency follows from by setting equal to the ion axial drift (in the perfectly neutral Lorentz frame). Three useful forms of this frequency are
(13) |
where is ion thermal velocity, is pinch radius, is ion inertial length, is characteristic ion cyclotron frequency, and the ion Budker parameter Crews, Meier, and Shumlak (2024).
Applying Eq. 13 to Eq. 12 and supposing , we infer three Lagrangian invariants associated to an ion fluid element threaded by betatron orbits, in analogy to Eqs. 5, to be
(14a) | |||
(14b) | |||
(14c) |
having assumed -isotropy, corresponding to the thermal axial invariant (Eq. 14a), radial invariant (Eq. 14b) and azimuthal angular momentum invariant (Eq. 14b). In Eqs. 14, is the characteristic flux density in the vicinity of the magnetic O-point about which the betatron orbits are bouncing. Combination of Eqs. 14 obtains the CGL model for an ion betatron fluid element,
(15a) | |||||
(15b) |
The electrical current axis assumes the principal anisotropy axis of the betatron fluid, and is otherwise modeled by the same CGL equations as a cyclotron plasma fluid. The radial and azimuthal directions display harmonic motions essentialized by a characteristic distance and a characteristic velocity (similar to the longitudinal mirror invariant), and the harmonic motion of the axial direction is essentialized by characteristic kinetic energy and the characteristic cyclotron frequency (like the magnetic moment of the cyclotrons). Anisotropic heating of the radial and axial directions is an intuitive consequence of the inductive axial electric field associated with flux compression splitting its energy between: 1) acceleration of the betatron axial drift, and 2) particle heating through increasing the energy of the betatron’s radial, axial, and azimuthal periodic motions.
Equations 14 come with important caveats. These invariants have been inferred based on the properties of “linear” betatron orbits, for which , which translates to a very small trapping parameter in the prequel’s terminology Crews, Meier, and Shumlak (2024). This analysis based on the linear orbit’s invariants parallels invariance of the magnetic moment for the cyclotron trajectories only to lowest order in the finite Larmor radius (FLR) parameter, viz. trapping parameter. The action integrals of a particular orbit are truly nonlinear functions of its trapping parameter.
Indeed, for the Bennett distribution, nonlinearity is non-negligible for many near-axis orbits. Sufficiently close to the magnetic O-point, these adiabatic invariants are expressed in elliptic integrals. Such nonlinear expressions for the radial betatron action were given explicitly by Sonnerup Sonnerup (1971) in the context of a reconnecting current sheet. Sonnerup’s analysis identified a significant fact: adiabatic compression of the center of the current sheet causes all orbits (both cyclotrons and betatrons) to approach the same elliptic modulus (trapping parameter) (at constant canonical momentum ). This trapping parameter represents a nonlinear betatron oscillation Crews, Meier, and Shumlak (2024). Thus, a corollary of Sonnerup’s result, originally identified for adiabatic compression of a reconnecting current sheet, is that sufficient flux compression of the Z pinch drives all near-axis particles into a nonlinear betatron orbit.
Sonnerup’s result assumes that the non-adiabatic cyclotron-betatron separatrix crossing is adiabatic-in-mean, as argued by Janicke Janicke (1975). In reality, the separatrix crossing is chaotic, as explored by Cary et al. Cary, Escande, and Tennyson (1986) and Büchner Büchner and Zelenyi (1989), which results in compression timescale-dependent corrections to the spectrum of the resulting betatron orbit flux. We note however that such non-adiabatic cross-separatrix orbital variations are less significant for the compressing Z pinch than the reconnecting current sheet because the infinite-period separatrix only occurs for purely radial/axial bounce orbits. Non-zero angular momentum of axis-encircling orbits erases the infinite-period separatrix, although all trajectories passing directly through may undergo these chaotic non-adiabatic changes. Non-adiabatic transitions across the betatron-cyclotron separatrix of axis-crossing orbits in a compressing Z pinch deserves deeper investigation in a future study.
III.3 Simple model of anisotropic heating from flux compression of a Z pinch
Sections III.1 and III.2 established that the CGL model applies to plasma composed of either cyclotron or betatron trajectories, in the usual way for the cyclotron plasma and a modified way for the betatron plasma. Specifically, the models predict the same magnitude of anisotropic heating from flux compression of a pinched plasma displaying either flavor of motion, but predict differing principal anisotropy axes; either the magnetic field lines for fluid consisting of cyclotron trajectories, or the electrical current lines for a fluid of betatron trajectories.
We now combine Sections III.1 and III.2 to propose an anisotropy model for a compressing Z pinch plasma, similar to the Ochs-Fisch model Ochs and Fisch (2018) but valid for all radii. The model consists of estimating a radially uniform magnitude of anisotropy due to flux compression using the CGL model and calculating the principal axis of the anisotropic Maxwellian thus generated in terms of the local partial pressures of cyclotron and betatron trajectories. Based on numerical experiments (Section X), this model appears accurate under the assumptions:
-
1.
The flux function varies slowly compared to the Alfvén transit time, and thus changes uniformly throughout the plasma;
-
2.
The change in particle magnetization during compression (due to separatrix crossing) is small enough that the relative densities of cyclotron and betatron trajectories are invariant.
Assumption 1 means that Eq. 6b may be used to estimate the magnitude of the generated anisotropy for all radii, and Assumption 2 that the anisotropy vector components in the final state may be parametrized using the initial state’s linear density.
Given Assumptions 1 and 2, the anisotropy model consists of the equations,
(16a) | |||||
(16b) | |||||
(16c) |
where and are the radially varying betatron and cyclotron orbital density fractions (cf. the prequel for the analytic fractions computed for the Bennett pinch Crews, Meier, and Shumlak (2024)). The anisotropies appearing in Eqs. 16 are defined as
(17a) | |||||
(17b) |
and are referred to as the agyrotropic and gyrotropic anisotropies in the following.
Figure 2 shows the gyrotropic and agyrotropic ion anisotropy profiles predicted by Eqs. 16 for various ion Budker parameters. The model predicts that agyrotropic anisotropy (induced by collisionless compression of ion betatron orbits) is dominant for small Budker parameters, while gyrotropic anisotropy (from collisionless compression of ion cyclotron trajectories) is dominant in the opposite limit. In the intermediate regime, , ion agyrotropy terminates in the vicinity of the pinch radius, which has important consequences for the transport of axial momentum and diamagnetic response of the compressing plasma discussed in the remainder of this work.

III.4 Additional sources of agyrotropic anisotropy in the Z pinch
In addition to adiabatic compression, agyrotropic anisotropy may be induced non-adiabatically due to collisionless shock heating Gedalin (2022) during fast Z pinch compression. Agyrotropy is induced in collisionless shock heating due to electric field gradients on the scale of the ion inertial length. In addition, radial electric field gradients induced by entropy mode fluctuations Ricci, Rogers, and Dorland (2006) and ensuing kinetic turbulence could also contribute to agyrotropy in ion distributions. Agyrotropic heating is frequently observed in simulations of kinetic plasma turbulence on ion inertial length scales Servidio et al. (2015); Yang et al. (2023). In this connection, note that species inertial length is directly proportional to pinch radius through the species Budker parameter Mahajan (1989); . Finally, strong electric fields can significantly modify particle orbits and drifts, particularly around magnetic O-points Joseph (2021).
III.5 A particular example of betatron heating
Figure 3 presents a visual aid to the discussion of Section III.2, presenting an example of betatron heating and acceleration in a compressing azimuthal magnetic flux. An ion with is evolved in the model time-dependent -directed vector potential , with and a specified flux compression function. The ion is released from with and , and the flux factor is linearly and slowly varied from an initial value to a final value.

The inductive electric field does work on the ion, increasing both the radial bounce energy and the axial drift/bounce energy in step (minus ). Most of the axial oscillation energy is channeled into the axial drift motion, increasing the axial bounce energy in the drifting frame much less than the radial bounce energy, because most of the axial energy in Fig. 3 is composed of the Galillean energy boost from the oscillation-center frame to the lab frame. Thus, betatron heating displays radial/axial temperature anisotropy because the change in radial bounce energy is distinct from the change in axial bounce energy in the drifting frame.
During collisionless adiabatic flux compression of a large Larmor radius Z pinch, mean ion betatron drift changes self-consistently with plasma current and density, and mean radial bounce energy varies so that radial pressure balances the self-magnetic energy (Bennett relation). Axial temperature is unconstrained by radial force balance, but is constrained by kinetic equilibrium considerations (which was not appreciated in early Z-pinch studies, see for example the claim in Bennett 1934 that axial temperature is irrelevant to equilibrium Bennett (1934)).
IV Kinetic equilibrium of magnetically self-focused flow
Collisionless kinetic equilibrium is a state where the distribution of particle velocities remains constant in time in the absence of collisions, unlike the fluid picture where collisions are assumed to be frequent enough to maintain a near-Maxwellian velocity distribution. Kinetic equilibria are stationary, self-consistent solutions to coupled kinetic-field equations, e.g., the Vlasov-Maxwell system. Collisionless plasmas find preferred equilibria through phase-mixing processes, meaning the dynamical phase-volume preserving rearragement of phase fluid into a lower energy state, i.e. waves and instabilitiesHelander and Mackenbach (2024). Most kinetic equilibria are not such minimum-energy states. Collisional kinetic equilibria, when relaxation and nonthermal forcing are in balance, are the starting point of all transport closures, i.e., the zero-order state of the plasma is non-Maxwellian.
Collisionless kinetic equilibrium is any function of the constants of motion (Jeans’ theorem). As discussed in the prequel Crews, Meier, and Shumlak (2024), for the flow pinch these constants are the energy and axial canonical momentum . The distribution is assumed to be canonical in the energy () due to phase mixing and quasineutrality. Recent work has shown that the maximum entropy state of the canonical ensemble is described by the family of kappa distributions Livadiotis and McComas (2013) which, although a universal description Livadiotis and McComas (2024), relies on a free parameter determined through a data fit. This would unnecessarily complicate the present analysis, which is taken to its limit in the Boltzmann distribution.
The primary nonthermal features of the flow Z pinch arise in the distribution’s dependence on canonical momentum. In this case, an approach which we call Suzuki’s Hermite method Suzuki and Shigeyama (2008) allows the construction of self-consistent kinetic equilibria using the Hermite polynomials associated with the canonical energy distribution. Suzuki’s Hermite method determines a generating function in the canonical momentum whose derivatives generate the fluid moments self-consistently satisfying the field equations, ensuring a physically valid equilibrium. An early study of the method is found in Channell Channell (1976), and a recent thorough study appears in Allanson et al Allanson et al. (2016).
We demonstrate the versatility of Suzuki’s Hermite method by presenting several examples of kinetic equilibria, including the well-known Bennett pinch equilibrium and a sheared-flow equilibrium with anisotropic pressure. The latter is particularly interesting as it provides a framework for understanding how velocity shear and temperature anisotropy interact in the collisionless regime. In Section V we describe this interaction in the context of collisionless gyroviscosity, which could play a crucial role in the stability and dynamics of sheared flow Z pinches. Here, we also explore the relationship between the generating function of Suzuki’s method and the plasma flow profile, revealing a deep connection between the kinetic and fluid descriptions of the Z pinch.
IV.1 Suzuki’s Hermite method of kinetic equilibrium
To begin, we present the aforementioned method Suzuki and Shigeyama (2008) to generate a self-consistent kinetic equilibrium using the properties of Hermite polynomials. Consider the ansatz distribution
(18) |
with the energy, the inverse temperature, the Hermite polynomials, the -component of magnetic potential, and the electric potential. Equation 18 is a kinetic equilibrium with provided that
(19) |
as found by substitution of Eq. 18 into the kinetic equation. Here the vector potential is written in lower case when it takes on the units of velocity, , and it is normalized to the thermal speed . Equation 19 is satisfied if each term of the series balances such that
(20) |
Now, the Hermite polynomials are the orthogonal family with the property . Substitution of the Hermite property into Eq. 20 shows that the coefficient functions must satisfy the recurrence relation
(21) |
Iteratively substituting Eq. 21 into Eq. 18 gives the series
(22) |
with the first Hermite coefficient. The particle density and current density follow as
(23a) | |||||
(23b) |
because are orthogonal to all other over the Gaussian weighting (). Equation 22 solves the kinetic equation, and its moments consistently solve the Vlasov-Maxwell system provided it satisfies the nonlinear Poisson equations
(24a) | |||||
(24b) |
Thus, the function is a generating function for the equilibrium, determining density, current density, and distribution function in the laboratory frame through its derivatives. Recalling that axial magnetic vector potential is magnetic flux per unit length (), this means that the generating function and all of its derivatives (the fluid moments) are flux functions. Notably, the toroidal flux function in kinetic tokamak equilibrium can be treated analogously through an azimuthal momentum () expansion Kaltsas et al. (2024). Intuitively, the generating function of a distribution canonical in the energy is expressed in Hermite series due to the properties of the Weierstrass transform, for which is the kernel Allanson et al. (2016). We now explore a few paradigmatic equilibria in terms of simple generating functions.
IV.2 Semi-canonical (shifted Maxwellian) equilibrium
The kinetic equilibrium of radially uniform axial velocity is the isothermal Bennett pinch, whose distribution is semi-canonical (namely, a shifted Maxwellian). The function is exponential in ,
(25) |
Repeated differentiation of Eq. 25 and application of the Hermite generating function,
(26a) | |||||
(26b) |
arrives at the shifted Maxwellian distribution,
(27) |
Both the Bennett pinch (cylindrical) and Harris sheet (Cartesian) are generated by Eq. 25 as they satisfy (under quasineutrality in the center-of-charge frame) the Liouville equation
(28) |
with the curvature constant Ceccherini et al. (2005). The cylindrical solution to Eq. 28 is the Bennett pinch’s vector potential, given by
(29) |
where is the characteristic pinch radius and is the self-magnetic flux per unit length ( is the total current enclosed to infinity). Regarding the generating function, it is useful that the following combination is unity,
(30) |
for each species , where is the linear density (particles per unit length) of species . The first factor is the species Budker parameter . For the Bennett pinch, the inverse Budker parameter is the drift Mach number for that species Crews, Meier, and Shumlak (2024),
(31a) | |||||
(31b) |
From Eq. 30, it follows from the generating function that the Bennett pinch particle and current densities are given by Meier and Shumlak (2021)
(32a) | |||||
(32b) |
In the isothermal Bennett kinetic equilibrium, both ions and electrons have uniform flow velocities, so their relative velocity is uniform too. As long as the relative velocity of two species is radially uniform then the magnetic potential will satisfy Eq. 28 and will be of the Bennett form, Eq. 29. The equilibrium has the important property that there is no radial electric field in the center-of-charge frame Gratreau (1978) where (assuming ).
IV.3 Sheared-flow, anisotropic (shifted bi-Maxwellian) equilibrium
The generator of the next simplest equilibrium is quadratic in the flux, namely for constants , . This equilibrium consists of a two-temperature bi-Maxwellian distribution with an agyrotropic pressure anisotropy supporting a radially sheared axial flow Del Sarto, Pegoraro, and Califano (2016). Anisotropic collisionless equilibria with sheared flows were explored by Mahajan and Hazeltine Mahajan and Hazeltine (2000), whose discussion we summarize here in the context of the Z pinch.
This sheared anisotropic equilibrium is the simplest kinetic equilibrium displaying the relationship between shear and anisotropy, enabling an understanding of how these two phenomena interact in the collisionless regime, which is relevant to Z pinches approaching fusion conditions. All sheared-flow collisionless kinetic equilibria are “inviscid” because they represent a stationary flow profile, which is a notable property. However, as will be discussed later, the collisionless magnetized dynamics should still be considered highly viscous in the sense that initially out-of-equilibrium flow profiles will quickly find a kinetic equilibrium supporting a steady flow.
Proceeding by direct calculation from an ansatz distribution is the simplest way to show existence of the equilibrium and determine its most fundamental properties. Suppose the distribution function is a two-temperature (bi-)Maxwellian whose first velocity moment is radially sheared (with species subscripts suppressed),
(33) |
Substituting Eq. 33 into the Vlasov equation leads to
(34) |
where , a shear stress drive term , and the distribution anisotropy is (for ). Taking the first radial moment of Eq. 34 yields
(35) |
giving the equilibrium radial force balance because . The distribution function supports this force balance in equilibrium (that is, with ) if the driver , namely when the anisotropy and shear profile are related by
(36) |
Having assumed a radially constant anisotropy , the axial species velocity is then given by , meaning that the equilibrium velocity profile is a linear flux function.
The bi-Maxwellian distribution is arrived at from the generating function using the completeness relation for Hermite polynomials Wiener (1988), as the derivatives of the generating function are themselves Hermite polynomials. Applying the completeness relation for the sum results in a bi-Maxwellian. The algebra is simpler to proceed as above, and as conducted in Mahajan and Hazeltine Mahajan and Hazeltine (2000), who ultimately find the bi-Maxwellian equilibrium for the sheared Bennett equilibrium (following section) to be
(37) |
where is a constant related to the anisotropy. Equation 37 is a nice closed form result for the distribution function whose sheared flow is a linear flux function. Higher-order dependence of the velocity on the flux function is explored in Section IV.4, but the series appear not to sum cleanly into exponentials as in Eq. 37.
IV.3.1 Sheared flow, anisotropic Bennett equilibrium
The self-consistent two-species equilibrium has been presented in Channon Channon and Coppins (2001). For reference, this solution is repeated here, and Channon’s integration constants considerably simplified. In general, sheared two-species flows lead to magnetic potentials of a more complex form than the Bennett profile’s (Eq. 29). Nevertheless, if the electron and ion axial shear flows are equal then their relative velocity remains radially uniform. This is consistent with the bi-Maxwellian sheared equilibrium provided that the electron and ion agyrotropic anisotropies balance as . For , the electrons are provided a comparatively gentle agyrotropy with . This balance of anisotropies together with quasineutrality leads to the Liouville equation for potential (Eq. 28) and the magnetic potential retains the Bennett form (Eq. 29).
The species densities, velocities, magnetic field, and radial electric field are given by
(38a) | |||||
(38b) | |||||
(38c) | |||||
(38d) | |||||
(38e) |
where is the ion Budker parameter and the ion agyrotropic anisotropy. As a generalized Bennett equilibrium, all the usual relations apply, provided that the radial temperature is used to define the thermal velocity in the Bennett relation and Eq. 31b.
IV.4 Higher-order kinetic equilibria of magnetically self-focused flows
Collisionless kinetic equilibrium of a sheared-flow self-pinched plasma is not necessarily a bi-Maxwellian distribution, which occurs for an isothermal pinch of radially constant temperature anisotropy in which flow velocity is a linear function of magnetic flux. Experimental and numerical simulation dynamics likely display both gyrotropic and agyrotropic anisotropies, due both to flow shear and to the asymmetric responses of cyclotron and betatron trajectories to changes in flux; hence general kinetic equilibria should display both axial and azimuthal temperature gradients. Non-isothermal kinetic equilibria are reserved to a future work.
But even considering the isothermal solutions, velocity need not be a linear flux function. Equilibria can be found as any polynomial function of flux. This section studies more general kinetic equilibria such that the axial velocity is some polynomial of the flux function, and studies the characteristic features of these solutions. First observe that the axial velocity is given by a logarithmic derivative of the generating function,
(39) |
Thus, each exponential polynomial is associated with a velocity profile as a polynomial in the magnetic potential,
(40) |
where are some coefficients. Equation 40 is a formal power series expansion of axial velocity about the pinch axis. With the velocity expressed as such a flux function (Eq. 40), we then seek the Hermite series coefficients to determine the equilibrium distribution function.
Substitution of Eq. 40 into Eq. 22 produces the distribution by the associated function of , whose Hermite series coefficients are the derivatives of . Let
(41) |
so that . Application of Faà di Bruno’s formula to Eq. 40 yields the Hermite coefficients as
(42) |
where is the n’th complete Bell polynomial Bell (1934). Thus, the semi-canonical distribution function associated with the postulated velocity profile of Eq. 40 is
(43) |
IV.4.1 Solution in multiple-variable Hermite polynomials
The coefficients of the Hermite series in Eq. 43 can be expressed recursively using the multiple-variable Hermite polynomials originally introduced by Charles Hermite and reintroduced by Dattoli et al Dattoli et al. (1994). We illustrate the general theory by considering a series for the axial velocity up to second order in the magnetic flux,
(44) |
which will involve the three-variable Hermite polynomials. Repeated differentiation of the generating function (Eq. 41) applied to Eq. 44 yields the formula
(45) |
where is the three-variable Hermite polynomial, generated by the recurrence relation Dattoli et al. (1994)
(46) |
with initial condition , yielding , , , etc. The usual single-variable Hermite polynomials are recovered with . Thus, the Hermite series for the function of canonical momentum corresponding to Eq. 44 is
(47) |
Higher-order vector potential dependence involves higher-order Hermite polynomials. That is, velocity as a cubic flux function involves four-variable Hermite polynomials, etc., generalizing the pattern of Eq. 46 in an obvious way.
In Eq. 47, the coefficients of the Hermite series are related to the Airy polynomials introduced by Torre for applications in beam optics Torre (2012) (by ). Equation 47 shows that nonthermality of the equilibrium where velocity is a quadratic flux function is described by a cross-series of Airy-like and Hermite polynomials. A closed form of this series is unknown to the authors.
The equilibrium was numerically investigated for quadratic and cubic velocity flux functions. These were found to induce skewness and kurtosis, respectively, in the distribution function. Figure 4 shows numerically computed distributions (in their drifting frames) computed for Fig. 4a) unsheared flow and Fig. 4b)-d) axial velocity as a linear, quadratic, and cubic flux functions, respectively. The coefficients were found using recursion relation for the multiple-variable Hermite polynomials, similar to Eq. 47.

These kinetic equilibrium solutions demonstrate that sheared flows are stationary in Z-pinch plasmas only if the distribution function has certain non-thermal features. If temperatures are isothermal and the distribution is canonical in the energy, though possibly anisotropic, then the equilibrium velocity profile is a flux function. First-order dependence on the flux is associated with pressure anisotropy (Fig. 4b), quadratic dependence with skewness (Fig. 4c), and so on. Regarding the heat flux, observe that a sum of Figs. 4b) and c) is a close match to the equilibrium distribution shown in Fig. 7 of Malara 2022 Malara et al. (2022). The interplay between velocity shear and pressure anisotropy has important consequences for collisionless viscosity in the flow pinch.
V Dynamical kinetic gyroviscosity: self-organization & self-generation of magnetically self-focused sheared flow
The previous section studied collisionless kinetic flow-pinch equilibria, highlighting a sheared-flow, anisotropic-pressure equilibrium which paradigmatically illustrates the interplay of sheared flows and pressure anisotropy. The existence of a sheared-flow equilibrium is surprising from the perspective of unmagnetized media. Collisionless unmagnetized flows, i.e. free molecular flows, are highly viscous; stationary, finite-pressure sheared flows do not exist in the large Knudson number limit because the flow-transverse components of trajectories are free to phase mix. The magnetic field enables sheared equilibria through a counteracting gyrophase mixing Gedalin (2015) which tends to reduce agyrotropic anisotropy by smoothing out variations with respect to gyrophase angle. These equilibrium solutions are valid regardless of orbit magnetization.
This section explores the collisionless limit through the interplay of flow shear and pressure anisotropy which regulates non-equilibrium flows through the magnetized pressure-strain interaction. This can be understood as the collisionless manifestation of gyroviscosity. Analysis of the stress tensor reveals that magnetized flows respond to shear stress by radial transport of axial momentum, as in the usual viscous process, but momentum transport is inhibited close to kinetic equilibrium. Thus axial flows self-organize as a consequence of collisionless gyroviscosity, which does not fully damp out flow shear but rather regulates it towards a stationary state, i.e. a magnetized sheared-flow kinetic equilibrium.
Shear stress results from two kinds of forcing: from a driven flow, or from a driven anisotropy. We term the former the forward process, in which viscosity mixes an initially thermal sheared flow into a steady anisotropic sheared flow, freezing flow into a flux function. The inverse process, namely generation of sheared flow, occurs from pressure anisotropy driven, for example, through flux compression. Self-generation of sheared flow by viscosity was, to our knowledge, first anticipated by Velikhov Velikhov (1964) and Haines Haines (1965) in the context of the dynamic -pinch. Sheared-flow generation in the compressing Z pinch was studied with a hybrid Vlasov-fluid model by Channon and colleagues Channon, Coppins, and Arber (1999); Channon et al. (2004). After discussing the forward process, we elaborate here on the flow-generating inverse process discussed by Velikhov, Haines, and Channon et al.
Only from the point-of-view of kinetic equilibria are collisionless cross-field flows inviscid. The kinetic dynamics may still be regarded as highly viscous in the sense that a distribution supporting out-of-equilibrium flow rapidly mixes towards a “frozen” equilibrium, thus organizing and regulating flow rather than fully dissipating it, but developing non-thermal features in the process. The dynamical perspective clarifies the physics of the response of a current-carrying plasma to pressure fluctuations.
Dynamical collisionless gyroviscosity is a reactive dissipation in analogy to reactive resistance in electrical circuits. In fact, dynamical gyroviscosity can induce velocity dispersion in one species and not another when ion and electron magnetizations are decoupled through the mass ratio. In this way, it will be seen that dynamical gyroviscosity produces reactive electrical resistance in response to sufficiently rapid pressure fluctuations, which is associated with spatially varying pressure anisotropies. The usual notion of dissipative gyroviscosity (i.e., non-reactive, or quasi-static) arises through collisional smoothing of non-thermal features in the small Hall parameter limit (with respect to the dynamical scales, cf. Section II).
V.1 Suppression of shear stress by pressure anisotropy in a magnetic field
It is intuitive to first consider the forward process, namely viscous damping of sheared flows. To begin, we extend the equilibrium analysis of Mahajan and Hazeltine by considering the evolution of an agryotropic distribution initially out-of-equilibrium, that is, when the equilibrium anisotropy condition is not met. Assuming initial radial force balance and introducing a collision operator , Eq. 34 can be written as
(48) |
instantaneously at the initial time . Given the bi-Maxwellian form of , the collisionless term on the RHS of Eq. 48 may be rewritten as derivatives of , giving
(49) |
and revealing it to be a diffusive operator, where is the radial thermal velocity. The mixed-derivatives term of Eq. 49 satisfies many of the features of a collision operator, such as conservation of particles, energy, and momentum, but drives shear stresses by not conserving the off-diagonal pressure tensor components .
Adopting a simple collision model such as the Lenard-Bernstein form shows the mixed derivatives to introduce off-diagonal terms to the velocity space diffusion matrix. The Fokker-Planck equation then has the paradigmatic form
(50a) | |||||
(50b) |
where is the relaxation rate and the shear stress drive frequency. The off-diagonal terms produce shearing in the velocity space (demonstrated in Fig. 5) as they are symmetric.

Shear stress generation by is observed by taking the moment for of Eq. 49,
(51) |
Assuming the distribution to relax on a timescale (or observing the dynamics a short time later), the solution to Eq. 51 can be estimated as where is a relaxed anisotropy. The radial viscous stress (from ) is then
(52) |
Ignoring the density gradient and assuming anisotropy is constant radially produces the axial momentum transport equation,
(53) |
where is the vector potential measured with units of velocity. Equation 53 demonstrates that the condition is just the equilibrium state around which dynamic processes operate.
V.2 Pressure-strain interaction and collisionless gyroviscosity
We wish to understand how a distribution might “find” this collisionless sheared-flow equilibrium from an initial non-equilibrium state. This section shows that an initially out-of-equilibrium distribution does not find equilibrium, but attains a decaying state with similar properties. This occurs as follows: the out-of-equilibrium state generates shear stresses (Eq. 51). These shear stresses produce pressure anisotropy and induce a viscous evolution of the plasma flow and current profiles. Shear stress production is inhibited (saturates) if the state is found. As observed by Del Sarto, residual shear stress in this state means that radial transport of axial momentum is mediated by magnetosonic wave emission Del Sarto, Pegoraro, and Tenerani (2017).
The discussion leading to Eq. 53 did not describe how pressure anisotropization is induced by Eqs. 50 because the assumed bi-Maxwellian distribution has zero shear stress. A more general theory is found by direct analysis of the Vlasov equation without an ansatz distribution. Del Sarto and Pegoraro performed such an analysis (neglecting heat fluxes) and derived an elegant matrix equation for the collisionless evolution of the ion pressure tensor Del Sarto and Pegoraro (2017),
(54) |
In Eq. 54, represent matrix commutators and anti-commutators. The matrix is the pressure tensor, the matrices and are the vorticity and (signed) cyclotron frequency dual tensors defined by and , the matrix is the incompressible rate of shear tensor, and is the isotropic compression. The matrix is the traceless symmetric part of the velocity gradient (strain rate) tensor.
We extend the discussion of Del Sarto and Pegoraro to analyze sheared flow about the axis of a magnetically self-focused flow by a direct calculation. Working in the -plane with sheared axial flow and taking the azimuthal angle to be an ignorable coordinate, the Cartesian theory is applicable provided that and the matrices are
(55a) | |||||
(55b) | |||||
(55c) |
Using Eqs. 55, the commutators work out to
(56a) | |||||
(56b) | |||||
(56c) |
where again is the agyrotropy coefficient defined previously, which are substituted into Eq. 54.
V.2.1 Viscous dissipation of collisionless unmagnetized flow
For intuition, consider first the collisionless unmagnetized case with (essentially free molecular flow), in which case
(57) |
The components of Eq. 57 show that temperature anisotropy develops during relaxation of collisionless sheared flow as the evolving shear stress drives diffusion through the viscous momentum flux , and concurrently axial temperature increases as . The anisotropy occurs as, in the collisionless limit of viscosity, heating occurs only in the flow direction. This axial heating is, of course, purely a consequence of the ballistic phase mixing associated with radial temperature. The process continues until and the shear stress asymptotically converges to a constant value. Fetsch and Fisch have recently noted that such anisotropic heating in fact increases fusion reactivity Fetsch and Fisch (2024).
V.2.2 Gyroviscous mixing of collisionless magnetized flow
The magnetic field introduces gyrophase mixing, tending to reduce the gyrophase variations induced by the axial heating due to mixing of nearby trajectories in Section V.2.1. With , Eq. 57 is instead,
(58) |
Consider the collisionless evolution of an initially isotropic sheared-flow Z pinch with . Shear stress develops from the drive term , but also produces pressure anisotropy through the component equations and , which combine into,
(59) |
Equation 59 shows that is the critical condition for the rate of viscous radial heating to exceed axial heating, thereby producing anisotropy and slowing the rate of shear stress production through . Physically, this indicates that the ion gyrofrequency exceeds the angular velocity of fluid element rotation in the sheared flow (). Shear stress production is minimized if the state is found, but equilibrium will not be found (as heating continues in the radial and transverse directions) without a dissipative process to remove the shear stress. This appears to produce a dynamic oscillation which pumps energy into the radial pressure, leading to sustained magnetosonic wave emission Del Sarto, Pegoraro, and Tenerani (2017).
In other words, inspection of Eq. 58 suggests that the kinetic equilibrium satisfying with (the bi-Maxwellian) is not an attractor as kinetic dynamics do not drive the plasma to a static state. However, the condition should be dynamically attained as it minimizes the time-evolving shear stresses through gyrophase mixing. A characteristic signature of this reorganization should be radial magnetosonic wave emission. Thus, collisionless gyroviscosity encourages fluid vorticity to be related to the gyrofrequency through phase mixing. In the isothermal limit, this makes the flow a flux function. Self-consistent Vlasov-Maxwell simulations are planned to better understand this nonlinear process.
V.3 Self-generated sheared axial flow: inverse gyroviscosity
Modifying the thought experiment of Section V.2.2 into an initially agyrotropic () but quasi-static (shear-free, ) plasma, we find that the inverse process of viscosity occurs: a sheared axial flow is spontaneously generated. Haines Haines (2011) (above their Eq. 3.33) has commented on the possibility of spontaneous sheared axial flow from finite Larmor orbit effects, and Channon Channon, Coppins, and Arber (1999); Channon et al. (2004) has observed the effect in hybrid Vlasov-fluid simulations of pinch compression. In this section, we thoroughly discuss the phenomenon of sheared flow generation via pressure anisotropy, as induced by adiabatic compression of betatron orbits in the case of Channon’s results.
The shear stress is estimated from Eq. 49 after a short time as . The axial component of the momentum equation now reads
(60) |
Equation 60 can be understood as an axial force arising by the gyroviscous stress of induced agyrotropic anisotropy. There is zero net force () as the shear stress decays much faster than , thus conserving total momentum like any viscosity.
The axial force density (Eq. 60) is approximated by taking the magnetic field and pressure to be Bennett distributed, and the agyrotropic anisotropy is modeled as due to betatron heating (Eq. 17a) and thus proportional to the density of betatron trajectories. Under these assumptions, the shear stress after time is proportional to the betatron fluid pressure and local cyclotron frequency,
(61) |
with the constant anisotropy magnitude (predicted by the CGL model), the local density of ion betatron trajectories, and the local ion cyclotron frequency. Equation 61 should be compared to the collisional closure with the total scalar pressure Miller and Shumlak (2016).
Figure 6 plots the anticipated axial force (from Eqs. 60 and 61) for various ion Budker parameters (Eq. 31a), showing the predicted oscillatory flow to reverse direction at the edge of the betatron density, or in other words towards the limit of the transitional magnetization region. This prediction agrees with the varying radial limit of sheared flow with large Larmor radius parameter (our Budker parameter) observed by Channon Channon et al. (2004). The generation of flow is nothing other than a viscous response to a pressure fluctuation, but is somewhat surprising at first glance as the plasma is initially static.
In our situation of a quasineutral large ion Larmor radius current-carrying plasma, essentially only the betatron ions experience this viscous response. The electrons mostly follow cyclotron trajectories and heat gyrotropically. It is interesting to observe that the force on the ions in response to the pressure change is counter the electric current. Thus, electrical current diffuses outwards. Electric current diffusion leads to magnetic diffusion, and hence is resistive. Thus, the self-generated ion flow resulting from flux compression illustrates an electrical resistance due purely to viscosity. This resistance is reactive in the sense that it arises from dynamic changes in the plasma. We suspect that the reactive gyroviscoresistivity described here may play a so-far underappreciated role as a dissipative process in dynamic Z pinches.

V.4 Spontaneous rotation from an axial magnetic field
Section V.3 observed that gyroviscosity drives spontaneous sheared axial flow due to agyrotropic anisotropy from e.g., collisionless betatron heating. One naturally wonders if similar phenomena arise from the gyrotropic anisotropy imbued to the cyclotron fluid. The answer is affirmative given a small frozen-in axial magnetic field Lucibello (2024a), in which case a spontaneous plasma rotation is predicted. Similarly to the betatron heating case, such a response is reactively gyroviscoresistive with respect to the azimuthal/poloidal currents. The differential rotation of a compressing -pinch is somewhat well-known and was first predicted by Velikhov Velikhov (1964). In the general picture, doubly adiabatic compression of a screw pinch leads to differential rotation within its magnetized periphery and differential axial flow within its current-dense core, as observed by Channon Channon et al. (2004).
Consider the pressure tensor dynamics with an axial magnetic field, assumed to be much smaller than the Z-pinch azimuthal flux. To be cautious in cylindrical coordinates, we directly derive the equation of Del Sarto and Pegoraro (Eq. 54) from the polar Vlasov equation Vogman, Shumlak, and Colella (2018), taking a separable ansatz that with a quasineutral plasma density and , azimuthal and axial flow velocities. The resulting kinetic equation is
(62) |
with and the cyclotron frequencies due to the - and -oriented magnetic fields. Taking moments for the pressure tensor () and neglecting all third and fourth-order moments (i.e., heat fluxes ) does obtain Del Sarto and Pegoraro’s result,
(63a) | |||||
(63b) | |||||
(63c) | |||||
(63d) | |||||
(63e) | |||||
(63f) |
with the gyrotropic anisotropy, which we model with Eq. 17b to occur typically far from the magnetic axis. The cylindrical terms (such as Coriolis force) do appear to play a role in heat flux dynamics. Given gyrotropic anisotropy, the axial magnetic field induces gyrotropic shear stress (Eq. 63e), inducing rotational force on the plasma in a manner similar to Eq. 60. As a diffusion, the spontaneous rotation is differential, reversing direction radially.
V.5 Characteristic anisotropy of Alfvénic sheared flow
Sheared flows are thought to decrease the growth rate of kink instabilities of flow pinches Shumlak and Hartman (1995), particularly in combination with viscoresistive effects Spies (1988) or axial magnetic flux Lucibello (2024b). It is thought that the flow shear should be comparable to the Alfvénic timescale for stability. Collisionless gyroviscosity phenomena both regulate and produce sheared flows in the collisionless large Larmor radius regime, i.e. when the ion inertial length is comparable to the pinch radius and the ion relaxation time is weakly comparable to the axial flow-through time or other dynamical timescales.
In the collisionless large Larmor radius regime, the ion velocity profile of a strongly forced flow viscously evolves to be proportional to the magnetic vector potential profile, self-organizing flow shear onto the scale of the pinch radius. The magnitude of shear is related to the magnitude of pressure anisotropy. In addition, anisotropy induced by doubly adiabatic flux compression generates radially sheared axial flow. These phenomena are potentially stabilizing to short-wavelength kink instabilities. The role of dissipation in the stability of such flows is under investigation.
It is thought that short-wavelength kink modes may be stabilized when the shear frequency is on the order of the Alfvén transit rate, . Considering leads to an estimate of the characteristic anisotropy associated with Alfvénic shear flow,
(64) |
where is the ion Budker parameter (Eq. 31a), equivalently the inverse drift parameter Crews, Meier, and Shumlak (2024) where . Thus, the characteristic anisotropy of a sheared-flow pinch in kinetic equilibrium is on the order of its drift parameter.
The ion Budker parameter depends only on the linear density of confined particles. Hence, a large number of confined particles (implying mostly ion cyclotron orbits, extensively discussed in the prequel Crews, Meier, and Shumlak (2024)) also requires only a gentle anisotropy for stabilizing self-organized shear flow in the collisionless regime. However, such anisotropy becomes hard to generate as the cyclotron orbits discourage agyrotropic anisotropy. Yet if the magnitude of ion anisotropy is too great then kinetic instabilities will enhance dissipation Bott, Cowley, and Schekochihin (2024). This suggests to target a moderate ion Budker parameter. To conclude, recent studies have noted the role of flow shear on the ion skin depth scale in regulating dissipation in astrophysical turbulence studies Hellinger, Petr et al. (2024). In the laboratory, the stabilizing effect of sheared flows on the tokamak’s internal kink have been observed to be concurrent with thermal anisotropy Bai et al. (2023).
VI Numerical validation of the anisotropy model
This section explores the validity of the anisotropy model, and concurrent shear flow generation, described analytically in this work. This is accomplished via a simple numerical experiment, namely evolving trajectories in a prescribed flux compression and observing the resulting statistics. The results support the hypothesis of self-generation of radially sheared axial flow due to doubly adiabatic compression of ion betatron orbits during ideal flux compression of a large Larmor radius Z pinch. The numerical experiment is a simple approximation to the plasma kinetics of flux compression, and is not self-consistent because neither the self-magnetic field of the generated ion flow nor its associated radial electric field are modeled. Nevertheless, this simple model provides insight into the essential physical models described in this article, validating the postulated anisotropy profiles and the axial flow generation mechanism. Self-consistent hybrid particle-in-cell simulations of sheared flow generation during large Larmor radius Z-pinch compression can be found in Channon et al Channon et al. (2004). A future study of the nonlinear dynamics of collisionless flux compression is planned to utilize self-consistent, continuum kinetic Vlasov-Maxwell simulations.
VI.1 Numerical model: evolving trajectories in a compressing flux function
The collisionless Boltzmann equation is a continuum description of the phase flow generated by an ensemble of discrete trajectories subject to a certain force field. A numerical solution of the collisionless Boltzmann equation is provided by sampling a large number of states from an initial distribution function, evolving the states in a specified field (from either a known solution or from expected dynamics), and reconstructing the distribution function of the evolved states. For example, kinetic equilibrium of the Bennett profile may be verified by sampling the initial distribution (an isothermal drifting Maxwellian with density radially distributed as ), evolving the trajectories in the Bennett field, and verifying stationary statistics of the observables. The method of observing statistics from trajectory ensembles is fairly common in turbulence studies Guillevic et al. (2023), and in this case is applied to observe development of pressure anisotropy and macroscopic flows resulting from flux compression of the Z pinch.
The method consists of the following. The center-of-charge frame is adopted (cf. the prequel Crews, Meier, and Shumlak (2024)) in which electrons and singly charged ions are perfectly counterstreaming and the electromagnetic field is entirely magnetic, generated by the two-parameter azimuthal flux per unit length
(65) |
where is the total current and characteristic radius. Suppose that Eq. 65 describes both the initial and final states of flux compression through the parameter sets and , connected by specified functions of time and representing time-varying current and geometry. The two parameters may be imagined to vary independently Crews et al. (2024), relating changes in flux to the current and inductance variations. The axial electric field arising from the varying flux is
(66) |
The two terms on the right-hand side are the parts and of the inductive electromotive force represented per unit length as the axial electric field. Observe that comparing the radial, compressive velocity to the radial guiding center of a cyclotron trajectory gives, with ,
(67) |
Equation 67 suggests that compression is geometrically self-similar (maintaining the form of Eq. 65) if the current is held constant and only inductance is varied. Thus, for this numerical experiment the parameter is smoothly varied as
(68) |
while holding the parameter constant. The proposed model is a quasi-static approximation which should be valid if the variations are slow relative to the Alfvén transit time. The similarity of the below results to the self-consistent simulations of Channon et al. Channon et al. (2004) suggests that a physical flux transport model is not necessary to validate the doubly adiabatic anisotropy model nor the physical mechanism of sheared-flow generation.
Ion trajectories sampled from the initial Bennett equilibrium are evolved in the time-varying electric and magnetic fields of Eqs. 65 and 66. A radial electric field is not included, meaning that the compression takes place consistently in the center-of-charge frame and the electric field associated with flow shear is neglected. The numerical experiment therefore consists of a specified compression of the flux function, an observed axial acceleration and radial compression of the plasma ions, and an implied axial acceleration and radial compression of the plasma electrons.
Let the dynamic variables be normalized to the initial thermal state,
(69a) | |||||
(69b) | |||||
(69c) | |||||
(69d) |
where are ion charge and mass and is the initial thermal velocity. In the normalized representation, the ion distribution function and the flux function are
(70a) | |||||
(70b) |
where is the partition function and is the ion Budker parameter. The following section presents numerical simulations at varying ion Budker parameter . The tildes are now dropped on the normalized variables.
VI.2 Simulated sheared flows at varying orbit magnetization
This section considers the sheared flows produced by three simulations at ion Budker parameters , representing the mostly betatron, mixed betatron/cyclotron, and mostly cyclotron regimes. The simulations are conducted by sampling ions from the initial equilibrium distribution of Eqs. 70 using the Metropolis algorithm and evolving their trajectories while compressing the scale parameter of the flux function from to according to Eq. 68 over a period of four normalized (Alfvén) times (confirmed by experimentation to be sufficiently long to observe adiabatic behavior). The distribution function is then reconstructed using radial kernel density estimation (KDE) Pierce and Kim (2023). The elementary RK45 time-stepping algorithm of the SciPy odeint routine was found to be sufficient to reconstruct the relevant statistics given the crudeness of the approximate solution. A highly resolved, fully self-consistent continuum kinetic simulation of the flux compression process is reserved to future work.
VI.2.1 Anticipated radial anisotropy profiles
The CGL-like anisotropy model developed in this work for the large Larmor radius Z pinch (Eqs. 16) predicts that flux compression induces an approximately radially uniform magnitude of anisotropy in which the principal anisotropy axis rotates from field-aligned in the magnetized periphery to current-aligned in the current-dense core. The anisotropy magnitude is a function of the electric current and geometric compression ratios. Given the geometric compression ratio and the fixed current , the anisotropy magnitude is predicted to be , and the anisotropy axis should rotate smoothly through the transitional magnetization layer whose location and thickness, depicted in Fig. 2, is a function of the ion Budker parameter.
VI.2.2 Simulation results
The profiles of temperature, density, anisotropy, and axial Mach number determined from moments of the estimated distribution function following compression, as obtained by the SciPy gaussian_kde method, are plotted in Fig. 7. A predicted axial Mach number is plotted alongside the axial Mach number, based on assuming the axial flow to locally display the equilibrium condition and is thus obtainable in quadrature as
(71) |
where the agyrotropic anisotropy is defined in Eq. 17. Agreement of the observed velocity profile with Eq. 71 would suggest the distribution post-compression is ideally gyromixed into the bi-Maxwellian kinetic equilibrium. The observed flow is seen to follow the general trend of Eq. 71 with a disagreement in magnitude, suggesting higher-order corrections in the kinetic equilibrium beyond the simple isothermal bi-Maxwellian solution.
General features of the solutions depicted in Fig. 7 are:
-
•
spatially varying anisotropy profiles and which sum approximately to a constant, consistent with the anisotropy orientation being proportional to the local density of betatron and cyclotron orbits as hypothesized (cf. Fig. 2);
-
•
radially sheared axial flows, depressed in the core and elevated at the edge, and terminating at the transition point between the betatron and cyclotron orbital densities;
-
•
isothermal radial temperature and non-isothermal axial and azimuthal temperatures;
-
•
depressed axial temperatures associated to radii with mostly betatron orbits, and excess azimuthal temperatures associated to radii with mostly cyclotron orbits.
It is tempting to associate the depressed axial temperatures around the betatron orbit density to the betatron orbits having greater susceptibility to axial acceleration by the axial electric field than the cyclotron orbits. However, observe that the cyclotron orbits at large radii display the fastest axial velocities following compression. As shown in Section V.3, this occurs because the viscous force induced by anisotropic betatron heating pushes the ion betatron population in the current-dense core into the -direction, and in the positive -direction throughout the magnetization transition layer, thus viscously transferring axial momentum to larger radii, accompanied by the emission of sonic waves. Thus, betatron acceleration in the axial electric field of the compressing azimuthal flux viscously produces faster flow at the edge than in the core. This counterintuitive result demonstrates the importance of the full spectrum of orbits, shedding light on the finite-radius process considered by B. A. Trubnikov in the singular limit Trubnikov (1992).
The numerical experiments suggest that higher Budker parameter cases obtain a greater pressure compression ratio. This discrepancy is a consequence of normalization to the initial states of the simulations. Although the flux compression ratio is the same for each simulation (evidenced by the consistent anisotropy ratios which develop), the absolute value of magnetic flux is greater for greater Budker parameters, resulting in higher relative voltages of compression and greater energy transfer. Note also that the betatron-rich plasmas do not compress geometrically, which occurs only when the drift motion follows the flux surfaces.

Figure 8 illustrates the reconstructed distribution function following compression marginalized on the azimuthal velocity (), providing an alternative perspective on the radial profiles demonstrated in Fig. 7. The full betatron regime () displays radial heating and an axial dispersion in velocities, the mixed betatron/cyclotron regime () shows a more even axial/radial heating with a diminished drift velocity, and the mainly cyclotron regime () shows agyrotropic anisotropy only close to the axis. The distribution functions appear close to an anisotropic Maxwellian, indicating that moment models with full pressure tensor dynamics Srinivasan and Hakim (2018) should be able to observe the self-generated axial shear flow phenomena explored in this work.

These simulations suggest, within the obvious limits of the non-self-consistent model, that axial flows within the pinch core may be anticipated during collisionless flux compression of a Z pinch plasma, potentially stabilizing the plasma to the kink instability. The nature of the self-generated sheared flow profile depends strongly on the distribution of cyclotron and betatron orbits, which is controlled by the initial linear density of the pinch prior to compression. An initial deuterium density around () appears ideal to generate a sonic sheared flow that terminates just past the pinch radius. Understanding the self-consistent, gyroviscoresistive evolution of the self-magnetic field due to the induced ion current and the consequences of microinstabilities triggered by anisotropy remains for future work.
VI.3 The role of thermal anisotropies in radial force balance
The pressure profiles observed to evolve away from the Bennett profile with the specified flux compression, which can be understood through analysis of the radial forces induced by the pressure anisotropy. Specifically, consider a three-temperature Maxwellian distribution of uniform radial temperature but radially varying, distinct axial and azimuthal temperatures and ,
(72) |
with the partition function. Substitution of Eq. 72 into the polar form of the stationary Vlasov equation (Eq. V.4) and computing the first radial moment yields the pressure balance,
(73) |
where is the gyrotropic pressure anisotropy. The terms are the temperature gradient forces Vold et al. (2018), while the thermodynamic force is an apparent force in polar coordinates describing an imbalance of the Coriolis and centrifugal forces acting on the population within a fluid element. Given the observed temperature gradients produced by anisotropy, these three thermal forces are directed inwards, which may flatten and even elevate the density distribution around the axis. Such features are observed in Fig. 7.
VII Summary of notable results and discussion
Sheared flows are well-recognized elements of high-performance fusion plasmas. Magnetized plasma viscosity, naturally an important consideration for fusion concepts utilizing sheared flows, is connected with thermal anisotropiesMahajan and Hazeltine (2000). In turn, anisotropies are themselves regulated by particle orbits through the statistics of adiabatic invariants and kinetic instabilities. Orbital magnetization and kinetics are thus key considerations for magnetized viscosity.
In the prequel, the spatial distributions of cyclotron and betatron trajectories were determined for a kinetic pinch equilibrium Crews, Meier, and Shumlak (2024). This work applied such concepts to kinetic equilibrium and dynamics. The anisotropy model of Chew, Goldberger, and Low (CGL), often called doubly adiabatic theory, was extended by taking into account both the cyclotron and betatron orbits. The CGL equations were found to accurately describe the magnitude of thermal anisotropy developed in a plasma fluid element, yet distinct anisotropy axes were found depending on the proportion of the trajectories. Specifically, betatron fluids align their anisotropy with the electric current. Thus, adiabatic variations in the betatron trajectories are associated with agyrotropic anisotropy.
Inviscid flows and collisionless trajectories are usually antonyms; the collisionless limit is highly viscous. On the other hand, collisionless, magnetized sheared flow kinetic equilibria exist in which gyro-phase mixing balances with the mixing of nearby trajectories. Agyrotropic anisotropy is associated with such flows because gyro-phase mixing smoothes out variations in the gyro-phase angle. However, magnetized flows are not inviscid. It is instructive to consider the initial-value problem of a magnetized flow out of equilibrium, in which case phase mixing rapidly brings it to equilibrium. In this sense, the flow is highly viscous, in which gyroviscosity organizes the flow rather than dissipating it entirely. It is interesting that gyro-phase mixing occurs regardless of whether the underlying orbits are cyclotrons or betatrons, but the requisite anisotropy is more naturally associated with the betatron orbits.
To understand the consequences of such agyrotropic anisotropies, a general class of sheared-flow kinetic pinch equilibrium was studied using the method of Hermite polynomials. The simplest equilibrium is isothermal and bi-Maxwellian, in which flow is a linear flux function and in which flow shear is maximum where local .
Higher-order equilibria were obtained by an expansion of flow in the vector potential. Essentially arbitrary equilibrium flows are obtained, yet higher-order flux dependence is associated with higher-moment features such as skewness, kurtosis, etc. Thus, higher-order flux dependence is associated with progressively more non-Maxwellian features, and in this way the linear flux function flow appears more robust to collisions that higher-order flows.
Kinetic dynamics were then considered in forced shear flows and in response to forced anisotropies. The former is the forward process of viscosity, in which the flow is phase-mixed to equilibrium, in effect freezing the velocity profile into a simple flux function. The latter is an inverse viscous process in which out-of-equilibrium thermal anisotropy self-generates sheared flow, for example during sufficiently fast adiabatic compression.
The doubly adiabatic model for the betatron fluid was validated by evolving ion trajectories of an initial Bennett pinch during flux compression. A radially uniform anisotropy magnitude results, but the anisotropy is gyrotropic in the magnetized periphery and agyrotropic in the current-dense core. Radially sheared axial flows occur through the transitional magnetization layer, in agreement with kinetic equilibrium solutions.
Substantial work remains to be done in Z-pinch kinetic theory. This article considers only velocity gradients and gyroviscosity; temperature gradients remain to be studied. Beyond equilibrium, the oscillation spectrum and kinetic stability of Z-pinch plasmas for Budker parameters in which the ion betatron population cannot be neglected is an open problem. Some work has been done, valid in the magnetized periphery, utilizing gyrokinetic models. It is possible that a reduced kinetic description amenable to linear analysis can be developed using the cyclotron-betatron splitting proposed in the prequel Crews, Meier, and Shumlak (2024). This approach is intended to be pursued in future work.
Acknowledgements.
The authors would like to thank A.E. Robson for valuable discussions on anomalous resistivity that have influenced the development of these ideas, A.D. Stepanov for extensive discussions on betatron heating, and J. Coughlin for discussions on gyroviscosity. The information, data, or work presented herein is based in part upon work supported by the National Science Foundation under Grant No. PHY-2108419.Author Declarations
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Daniel W. Crews: Conceptualization (lead), Formal analysis (lead), Investigation (lead), Methodology (lead), Writing - original draft (lead), Writing - review and editing (lead). Eric T. Meier: Conceptualization (equal), Project Administration (equal), Supervision (equal), Writing - review & editing (equal). Uri Shumlak: Conceptualization (equal), Project Administration (equal), Supervision (equal), Writing - review & editing (equal).
Data Availability
The data that support the findings of this study are available from the corresponding author upon reasonable request.
References
- Bennett (1934) W. H. Bennett, “Magnetically self-focussing streams,” Phys. Rev. 45, 890–897 (1934).
- Tonks (1937) L. Tonks, “Theory and phenomena of high current densities in low pressure arcs,” Transactions of The Electrochemical Society 72, 167 (1937).
- Artsimovich (1962) L. A. Artsimovich, “Controlled Nuclear Fusion Research, September 1961: Review of experimental results,” Nuclear Fusion, Suppl. (1962).
- Bostick (1977) W. H. Bostick, “The pinch effect revisited,” International Journal of Fusion Energy 1, 1–55 (1977).
- Lawson (1958) J. D. Lawson, “Perveance and the Bennett Pinch Relation in Partially-Neutralized Electron Beams,” Journal of Electronics and Control 5, 146–151 (1958), https://doi.org/10.1080/00207215808953898 .
- Budker (1956) G. I. Budker, “Relativistic stabilized electron beam,” The Soviet Journal of Atomic Energy 1, 673–686 (1956).
- Linhart (1959) J. G. Linhart, “Accelerated Self-Constricted Electron Streams in Plasma,” Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences 249, 318–334 (1959).
- Alfvén (1939) H. Alfvén, “On the Motion of Cosmic Rays in Interstellar Space,” Phys. Rev. 55, 425–429 (1939).
- Lawson (1959) J. D. Lawson, “On the classification of electron streams,” Journal of Nuclear Energy. Part C, Plasma Physics, Accelerators, Thermonuclear Research 1, 31 (1959).
- Crews, Meier, and Shumlak (2024) D. W. Crews, E. T. Meier, and U. Shumlak, “Z Pinch Kinetics I - A Particle Perspective: Transitional Magnetization and Cyclotron and Betatron Orbits,” In preparation (2024).
- Arnold (2013) V. I. Arnold, Mathematical Methods of Classical Mechanics, Graduate Texts in Mathematics (Springer New York, 2013).
- Finkelstein and Sturrock (1959) D. Finkelstein and P. A. Sturrock, “Stability of relativistic self-focusing streams,” in Plasma Physics, Ed. J. E. Drummond (Dover, 1959).
- Bennett (1955) W. H. Bennett, “Self-focusing streams,” Phys. Rev. 98, 1584–1593 (1955).
- Haines (2001) M. G. Haines, “Kinetic effects in z pinches,” Laser and Particle Beams 19, 345–353 (2001).
- Santos et al. (2018) J. J. Santos, M. Bailly-Grandvaux, M. Ehret, A. V. Arefiev, D. Batani, F. N. Beg, A. Calisti, S. Ferri, R. Florido, P. Forestier-Colleoni, S. Fujioka, M. A. Gigosos, L. Giuffrida, L. Gremillet, J. J. Honrubia, S. Kojima, P. Korneev, K. F. F. Law, J.-R. Marquès, A. Morace, C. Mossé, O. Peyrusse, S. Rose, M. Roth, S. Sakata, G. Schaumann, F. Suzuki-Vidal, V. T. Tikhonchuk, T. Toncian, N. Woolsey, and Z. Zhang, “Laser-driven strong magnetostatic fields with applications to charged beam transport and magnetized high energy-density physics,” Physics of Plasmas 25, 056705 (2018), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.5018735/14700567/056705_1_online.pdf .
- Arefiev et al. (2024) A. Arefiev, I.-L. Yeh, K. Tangtartharakul, and L. Willingale, “Electron energy gain due to a laser frequency modulation experienced by electron during betatron motion,” Physics of Plasmas 31, 023106 (2024), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0190559/19725041/023106_1_5.0190559.pdf .
- Vikhrev and Korolev (2007) V. V. Vikhrev and V. D. Korolev, “Neutron generation from Z-pinches,” Plasma Physics Reports 33, 356–380 (2007).
- Goyon et al. (2024) C. Goyon, S. C. Bott-Suzuki, A. E. Youmans, J. T. Banasek, L. A. Morton, B. Levitt, J. R. Barhydt, K. D. Morgan, C. Liekhus-Schmaltz, W. C. Young, D. P. Higginson, A. C. Hossack, E. T. Meier, B. A. Nelson, M. Quinley, A. Taylor, P. Tsai, N. van Rossum, A. Shah, A. D. Stepanov, D. A. Sutherland, T. R. Weber, U. Shumlak, and H. S. McLean, “Plasma pressure profiles in a sheared-flow-stabilized Z-pinch,” Physics of Plasmas 31, 072503 (2024), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0209351/20027284/072503_1_5.0209351.pdf .
- Mitrani et al. (2021) J. M. Mitrani, J. A. Brown, B. L. Goldblum, T. A. Laplace, E. L. Claveau, Z. T. Draper, E. G. Forbes, R. P. Golingo, H. S. McLean, B. A. Nelson, U. Shumlak, A. Stepanov, T. R. Weber, Y. Zhang, and D. P. Higginson, “Thermonuclear neutron emission from a sheared-flow stabilized Z-pinch,” Physics of Plasmas 28, 112509 (2021), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0066257/13408570/112509_1_online.pdf .
- Kotelnikov (2020) I. Kotelnikov, “On the structure of the boundary layer in a Beklemishev diamagnetic bubble,” Plasma Physics and Controlled Fusion 62, 075002 (2020).
- Timofeev, Kurshakov, and Berendeev (2024) I. V. Timofeev, V. A. Kurshakov, and E. A. Berendeev, “Formation of cylindrical plasma equilibria with ,” Physics of Plasmas 31, 082512 (2024), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0216073/20127415/082512_1_5.0216073.pdf .
- Braginskii (1965) S. I. Braginskii, “Transport processes in a plasma,” in Reviews of Plasma Physics, Vol. 1 (Consulants Bureau, New York, 1965).
- Du et al. (2020) S. Du, G. P. Zank, X. Li, and F. Guo, “Energy dissipation and entropy in collisionless plasma,” Phys. Rev. E 101, 033208 (2020).
- Bandyopadhyay et al. (2023) R. Bandyopadhyay, Y. Yang, W. H. Matthaeus, T. N. Parashar, V. Roytershteyn, A. Chasapis, D. J. Gershman, B. L. Giles, and J. L. Burch, “Collisional-like dissipation in collisionless plasmas,” Physics of Plasmas 30, 080702 (2023), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0146986/18086698/080702_1_5.0146986.pdf .
- Del Sarto, Pegoraro, and Califano (2016) D. Del Sarto, F. Pegoraro, and F. Califano, “Pressure anisotropy and small spatial scales induced by velocity shear,” Phys. Rev. E 93, 053203 (2016).
- Del Sarto and Pegoraro (2017) D. Del Sarto and F. Pegoraro, “Shear-induced pressure anisotropization and correlation with fluid vorticity in a low collisionality plasma,” Monthly Notices of the Royal Astronomical Society 475, 181–192 (2017), https://academic.oup.com/mnras/article-pdf/475/1/181/23388297/stx3083.pdf .
- Yang et al. (2017) Y. Yang, W. H. Matthaeus, T. N. Parashar, C. C. Haggerty, V. Roytershteyn, W. Daughton, M. Wan, Y. Shi, and S. Chen, “Energy transfer, pressure tensor, and heating of kinetic plasma,” Physics of Plasmas 24, 072306 (2017), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.4990421/15999836/072306_1_online.pdf .
- Kaufman (1960) A. N. Kaufman, “Plasma Viscosity in a Magnetic Field,” Physics of Fluids (1960), 10.1063/1.1706096.
- Hellinger, Petr et al. (2024) Hellinger, Petr, Verdini, Andrea, Montagud-Camps, Victor, Franci, Luca, Papini, Emanuele, Matteini, Lorenzo, and Landi, Simone, “Anisotropy of plasma turbulence at ion scales: Hall and pressure–strain effects,” A&A 684, A120 (2024).
- Bandyopadhyay et al. (2020) R. Bandyopadhyay, W. H. Matthaeus, T. N. Parashar, Y. Yang, A. Chasapis, B. L. Giles, D. J. Gershman, C. J. Pollock, C. T. Russell, R. J. Strangeway, R. B. Torbert, T. E. Moore, and J. L. Burch, “Statistics of Kinetic Dissipation in the Earth’s Magnetosheath: MMS Observations,” Phys. Rev. Lett. 124, 255101 (2020).
- Pezzi et al. (2019) O. Pezzi, Y. Yang, F. Valentini, S. Servidio, A. Chasapis, W. H. Matthaeus, and P. Veltri, “Energy conversion in turbulent weakly collisional plasmas: Eulerian hybrid Vlasov-Maxwell simulations,” Physics of Plasmas 26, 072301 (2019), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.5100125/13820638/072301_1_online.pdf .
- Cassak and Barbhuiya (2022) P. A. Cassak and M. H. Barbhuiya, “Pressure–strain interaction. I. On compression, deformation, and implications for Pi-D,” Physics of Plasmas 29, 122306 (2022), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0125248/19799503/122306_1_online.pdf .
- Cassak, Barbhuiya, and Weldon (2022) P. A. Cassak, M. H. Barbhuiya, and H. A. Weldon, “Pressure–strain interaction. II. Decomposition in magnetic field-aligned coordinates,” Physics of Plasmas 29, 122307 (2022), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0125252/19799488/122307_1_online.pdf .
- Barbhuiya and Cassak (2022) M. H. Barbhuiya and P. A. Cassak, “Pressure–strain interaction. III. Particle-in-cell simulations of magnetic reconnection,” Physics of Plasmas 29, 122308 (2022), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0125256/19799538/122308_1_online.pdf .
- Conley et al. (2024) S. A. Conley, J. Juno, J. M. TenBarge, M. H. Barbhuiya, P. A. Cassak, G. G. Howes, and E. Lichko, “The kinetic analogue of the pressure-strain interaction,” (2024), arXiv:2408.06508 [physics.plasm-ph] .
- Shumlak (2020) U. Shumlak, “Z-pinch fusion,” Journal of Applied Physics 127, 200901 (2020), https://pubs.aip.org/aip/jap/article-pdf/doi/10.1063/5.0004228/19782092/200901_1_online.pdf .
- Ryutov, Derzon, and Matzen (2000) D. D. Ryutov, M. S. Derzon, and M. K. Matzen, “The physics of fast pinches,” Rev. Mod. Phys. 72, 167–223 (2000).
- Morozov and Solov’ev (1980) A. I. Morozov and L. S. Solov’ev, “Steady-state plasma flow in a magnetic field,” in Reviews of Plasma Physics, edited by M. A. Leontovich (Springer US, Boston, MA, 1980) pp. 1–103.
- Turchi and Shumlak (2023) P. J. Turchi and U. Shumlak, “Some Considerations of a Flow-Stabilized Z-Pinch for Megagauss Fusion,” IEEE Transactions on Plasma Science 51, 3679–3687 (2023).
- Crews et al. (2024) D. W. Crews, I. A. M. Datta, E. T. Meier, and U. Shumlak, “The Kadomtsev Pinch Revisited for Sheared-Flow-Stabilized Z-Pinch Modeling,” IEEE Transactions on Plasma Science , 1–13 (2024).
- Hunana et al. (2022) P. Hunana, T. Passot, E. Khomenko, D. Martínez-Gómez, M. Collados, A. Tenerani, G. P. Zank, Y. Maneva, M. L. Goldstein, and G. M. Webb, “Generalized Fluid Models of the Braginskii Type,” The Astrophysical Journal Supplement Series 260, 26 (2022).
- Meier and Shumlak (2021) E. T. Meier and U. Shumlak, “Development of five-moment two-fluid modeling for Z-pinch physics,” Physics of Plasmas 28, 092512 (2021), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0058420/13420598/092512_1_online.pdf .
- Ryutov (2015) D. D. Ryutov, “Characterizing the Plasmas of Dense -Pinches,” IEEE Transactions on Plasma Science 43, 2363–2384 (2015).
- Shumlak, Meier, and Levitt (2024) U. Shumlak, E. T. Meier, and B. J. Levitt, “Fusion Gain and Triple Product for the Sheared-Flow-Stabilized Z Pinch,” Fusion Science and Technology 80, 1–16 (2024), https://doi.org/10.1080/15361055.2023.2198049 .
- Bott, Cowley, and Schekochihin (2024) A. F. A. Bott, S. C. Cowley, and A. A. Schekochihin, “Kinetic stability of Chapman–Enskog plasmas,” Journal of Plasma Physics 90, 975900207 (2024).
- Braginskii (1958) S. I. Braginskii, “Transport processes in a plasma,” Reviews of Plasma Physics 1, 358–369 (1958).
- Chew, Goldberger, and Low (1956) G. F. Chew, M. L. Goldberger, and F. E. Low, “The Boltzmann Equation and the One-Fluid Hydromagnetic Equations in the Absence of Particle Collisions,” Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences 236, 112–118 (1956).
- Sonnerup (1971) B. U. O. Sonnerup, “Adiabatic particle orbits in a magnetic null sheet,” Journal of Geophysical Research (1896-1977) 76, 8211–8222 (1971), https://agupubs.onlinelibrary.wiley.com/doi/pdf/10.1029/JA076i034p08211 .
- Janicke (1975) L. Janicke, “Adiabatische invarianz im mittel / adiabatic invariance in the average,” Zeitschrift für Naturforschung A 30, 421–430 (1975).
- Cary, Escande, and Tennyson (1986) J. R. Cary, D. F. Escande, and J. L. Tennyson, “Adiabatic-invariant change due to separatrix crossing,” Phys. Rev. A 34, 4256–4275 (1986).
- Büchner and Zelenyi (1989) J. Büchner and L. M. Zelenyi, “Regular and chaotic charged particle motion in magnetotaillike field reversals: 1. Basic theory of trapped motion,” Journal of Geophysical Research: Space Physics 94, 11821–11842 (1989), https://agupubs.onlinelibrary.wiley.com/doi/pdf/10.1029/JA094iA09p11821 .
- Ochs and Fisch (2018) I. E. Ochs and N. J. Fisch, “Anisotropy-driven collisional separation of impurities in magnetized compressing and expanding cylindrical plasmas,” Physics of Plasmas 25, 122306 (2018), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.5055568/15779257/122306_1_online.pdf .
- Gedalin (2022) M. Gedalin, “Combining Rankine–Hugoniot relations, ion dynamics in the shock front, and the cross-shock potential,” Physics of Plasmas 29, 112904 (2022), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0120578/16630179/112904_1_online.pdf .
- Ricci, Rogers, and Dorland (2006) P. Ricci, B. N. Rogers, and W. Dorland, “Small-Scale Turbulence in a Closed-Field-Line Geometry,” Phys. Rev. Lett. 97, 245001 (2006).
- Servidio et al. (2015) S. Servidio, F. Valentini, D. Perrone, A. Greco, F. Califano, W. H. Matthaeus, and P. Veltri, “A kinetic model of plasma turbulence,” Journal of Plasma Physics 81, 325810107 (2015).
- Yang et al. (2023) Y. Yang, F. Pecora, W. H. Matthaeus, S. Roy, M. E. Cuesta, A. Chasapis, T. Parashar, R. Bandyopadhyay, D. J. Gershman, B. L. Giles, and J. L. Burch, “Quantifying the Agyrotropy of Proton and Electron Heating in Turbulent Plasmas,” The Astrophysical Journal 944, 148 (2023).
- Mahajan (1989) S. M. Mahajan, “Exact and almost exact solutions to the Vlasov–Maxwell system,” Physics of Fluids B: Plasma Physics 1, 43–54 (1989).
- Joseph (2021) I. Joseph, “Guiding center and gyrokinetic orbit theory for large electric field gradients and strong shear flows,” Physics of Plasmas 28, 042102 (2021), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0037889/16004149/042102_1_online.pdf .
- Helander and Mackenbach (2024) P. Helander and R. J. J. Mackenbach, “Available energy of plasmas with small fluctuations,” Journal of Plasma Physics 90, 905900401 (2024).
- Livadiotis and McComas (2013) G. Livadiotis and D. J. McComas, “Understanding kappa distributions: A toolbox for space science and astrophysics,” Space Science Reviews 175, 183–214 (2013).
- Livadiotis and McComas (2024) G. Livadiotis and D. J. McComas, “Universality of kappa distributions,” Europhysics Letters 146, 41003 (2024).
- Suzuki and Shigeyama (2008) A. Suzuki and T. Shigeyama, “A novel method to construct stationary solutions of the Vlasov–Maxwell system,” Physics of Plasmas 15 (2008).
- Channell (1976) P. J. Channell, “Exact Vlasov–Maxwell equilibria with sheared magnetic fields,” The Physics of Fluids 19, 1541–1545 (1976), https://pubs.aip.org/aip/pfl/article-pdf/19/10/1541/12443456/1541_1_online.pdf .
- Allanson et al. (2016) O. Allanson, T. Neukirch, S. Troscheit, and F. Wilson, “From one-dimensional fields to Vlasov equilibria: theory and application of Hermite polynomials,” Journal of Plasma Physics 82, 905820306 (2016).
- Kaltsas et al. (2024) D. A. Kaltsas, A. Kuiroukidis, P. J. Morrison, and G. N. Throumoulopoulos, “Axisymmetric hybrid Vlasov equilibria with applications to tokamak plasmas,” Plasma Physics and Controlled Fusion 66, 065016 (2024).
- Ceccherini et al. (2005) F. Ceccherini, C. Montagna, F. Pegoraro, and G. Cicogna, “Two-dimensional Harris–Liouville plasma kinetic equilibria,” Physics of Plasmas 12, 052506 (2005), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.1899083/13506459/052506_1_online.pdf .
- Gratreau (1978) P. Gratreau, “Generalized Bennett equilibria and particle orbit analysis of plasma columns carrying ultra-high currents,” The Physics of Fluids 21, 1302–1311 (1978), https://pubs.aip.org/aip/pfl/article-pdf/21/8/1302/12259969/1302_1_online.pdf .
- Mahajan and Hazeltine (2000) S. M. Mahajan and R. D. Hazeltine, “Sheared-flow generalization of the Harris sheet,” Physics of Plasmas 7, 1287–1293 (2000), https://pubs.aip.org/aip/pop/article-pdf/7/4/1287/19322230/1287_1_online.pdf .
- Wiener (1988) N. Wiener, The Fourier integral and certain of its applications (CUP Archive, 1988).
- Channon and Coppins (2001) S. W. Channon and M. Coppins, “Z-pinch Vlasov-fluid equilibria including sheared-axial flow,” Journal of Plasma Physics 66, 337–347 (2001).
- Bell (1934) E. T. Bell, “Exponential Polynomials,” Annals of Mathematics 35, 258–277 (1934).
- Dattoli et al. (1994) G. Dattoli, C. Chiccoli, S. Lorenzutta, G. Maino, and A. Torre, “Theory of generalized Hermite polynomials,” Computers and Mathematics with Applications 28, 71–83 (1994).
- Torre (2012) A. Torre, “Airy polynomials, three-variable Hermite polynomials and the paraxial wave equation,” Journal of Optics 14, 045704 (2012).
- Malara et al. (2022) F. Malara, A. Settino, D. Perrone, O. Pezzi, G. Guzzi, and F. Valentini, “Exact shearing flow magnetized hybrid kinetic equilibria with inhomogeneous temperature,” The Astrophysical Journal 941, 201 (2022).
- Gedalin (2015) M. Gedalin, “Collisionless relaxation of non-gyrotropic downstream ion distributions: dependence on shock parameters,” Journal of Plasma Physics 81, 905810603 (2015).
- Velikhov (1964) E. P. Velikhov, “A note on plasma rotation,” Journal of Nuclear Energy. Part C, Plasma Physics, Accelerators, Thermonuclear Research 6, 203 (1964).
- Haines (1965) M. G. Haines, “On the origin of rotation in magnetically confined plasmas,” Advances in Physics 14, 167–211 (1965), https://doi.org/10.1080/00018736500101031 .
- Channon, Coppins, and Arber (1999) S. W. Channon, M. Coppins, and T. D. Arber, “One-dimensional hybrid simulations of the imploding large Larmor radius Z pinch,” Physics of Plasmas 6, 3890–3897 (1999), https://pubs.aip.org/aip/pop/article-pdf/6/10/3890/19159658/3890_1_online.pdf .
- Channon et al. (2004) S. Channon, I. Mitchell, E. Wyndam, M. Favre, and H. Chuaquai, “Sheared Flow Generation in Imploding Vlasov-Fluid Plasmas,” Physica Scripta 69, 30 (2004).
- Del Sarto, Pegoraro, and Tenerani (2017) D. Del Sarto, F. Pegoraro, and A. Tenerani, “‘Magneto-elastic’ waves in an anisotropic magnetised plasma,” Plasma Physics and Controlled Fusion 59, 045002 (2017).
- Fetsch and Fisch (2024) H. Fetsch and N. J. Fisch, “Enhancement to fusion reactivity in sheared flows,” (2024), arXiv:2410.03590 [physics.plasm-ph] .
- Haines (2011) M. G. Haines, “A review of the dense Z-pinch,” Plasma Physics and Controlled Fusion 53, 093001 (2011).
- Miller and Shumlak (2016) S. T. Miller and U. Shumlak, “A multi-species 13-moment model for moderately collisional plasmas,” Physics of Plasmas 23, 082303 (2016), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.4960041/16032187/082303_1_online.pdf .
- Lucibello (2024a) P. Lucibello, “Z-Pinch Stabilization by a Frozen-In Axial Magnetic Flux,” IEEE Transactions on Plasma Science 52, 310–318 (2024a).
- Vogman, Shumlak, and Colella (2018) G. Vogman, U. Shumlak, and P. Colella, “Conservative fourth-order finite-volume Vlasov–Poisson solver for axisymmetric plasmas in cylindrical phase space coordinates,” Journal of Computational Physics 373, 877–899 (2018).
- Shumlak and Hartman (1995) U. Shumlak and C. W. Hartman, “Sheared Flow Stabilization of the Kink Mode in Pinches,” Phys. Rev. Lett. 75, 3285–3288 (1995).
- Spies (1988) G. O. Spies, “Visco-resistive stabilization of kinks with short wavelengths along an elliptic magnetic stagnation line,” Plasma Physics and Controlled Fusion 30, 1025 (1988).
- Lucibello (2024b) P. Lucibello, “On axial current induction and stability of a pinch,” IEEE Transactions on Plasma Science , 1–15 (2024b).
- Bai et al. (2023) X. Bai, Y. Liu, G. Hao, G. Dong, N. Zhang, and S. Wang, “A new mechanism for internal kink stabilization by plasma flow and anisotropic thermal transport,” Nuclear Fusion 64, 014001 (2023).
- Guillevic et al. (2023) A. Guillevic, M. Lesur, X. Garbet, P. Diamond, G. Lo-Cascio, Y. Kosuga, E. Gravier, D. Mandal, A. Ghizzo, and T. Réveillé, “Particle dynamics in a turbulent electric field,” Physics of Plasmas 30, 052301 (2023), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/5.0135918/17177512/052301_1_5.0135918.pdf .
- Pierce and Kim (2023) C. M. Pierce and Y.-K. Kim, “Kernel density estimators for axisymmetric particle beams,” Instruments 7 (2023), 10.3390/instruments7040044.
- Trubnikov (1992) B. A. Trubnikov, “A new hypothesis of cosmic ray generation in plasma pinches,” IEEE Transactions on Plasma Science 20, 898–904 (1992).
- Srinivasan and Hakim (2018) B. Srinivasan and A. Hakim, “Role of electron inertia and electron/ion finite Larmor radius effects in low-beta, magneto-Rayleigh-Taylor instability,” Physics of Plasmas 25, 092108 (2018), https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.5046098/15884196/092108_1_online.pdf .
- Vold et al. (2018) E. Vold, G. Kagan, A. N. Simakov, K. Molvig, and L. Yin, “Self-similar solutions for multi-species plasma mixing by gradient driven transport,” Plasma Physics and Controlled Fusion 60, 054010 (2018).