A&A 485, 327–336 (2008) Astronomy DOI: 10.1051/0004-6361:20078911 & c ESO 2008 Astrophysics

A Boltzmann-kinetical description of an MHD shock with arbitrary field inclination

M. Siewert and H.-J. Fahr

Argelander Institut für Astronomie der Universität Bonn, Abteilung f. Astrophysik und Extraterrestrische Forschung, Auf dem Huegel 71, 53121 Bonn, Germany e-mail: [email protected] Received 24 October 2007 / Accepted 27 March 2008

ABSTRACT

Aims. We revisit the general problem of the anisotropic MHD shock for arbitrary magnetic field inclinations, where the jump condi- tions are underdetermined. To describe the transition region of the shock, we derive a variant of a kinetic Boltzmann-Vlasov equation previously used to describe the perpendicular shock in the absence of dissipative processes. Methods. We derive effective force terms, for the kinetic equation, that are based on the conservation of the Chew-Goldberger-Low (CGL) MHD invariants which appear in the standard model for anisotropic MHD. This approach is based on a generalisation of the well-known equivalence between the first CGL invariant and the integral over the magnetic moments of the underlying particles. Results. Assuming an arbitrary distribution function on the upstream side, we integrate the kinetic equation across the shock. This result allows us to establish further relations between the MHD velocity moments on both sides. Using this additional information, we close the anisotropic MHD jump conditions. In addition, the now unique solution of the jump conditions allows us to present explicit cuts through a representative Maxwellian distribution function on both sides of the shock. In the kinetic equation, one only requires two parameters that need to be derived from the classical jump conditions, the classical MHD compression ratio and an equivalent ratio for the magnetic field strengths. Key words. plasmas – shock waves – magnetohydrodynamis (MHD) – sun: solar wind

1. Introduction (see e.g. Cercignani 1988), which needs to be truncated at an arbitrary point. We published previously several kinetic studies of an The second approach to this problem is based on experi- crossing an MHD shock, such as the bow shock of the Earth or mental observations, followed by appropriate modelling of these the solar wind termination shock (Fahr & Siewert 2006; Siewert data. The cluster mission (Escoubet et al. 1997) has observed & Fahr 2007a,b), that attempted to improve our understanding of the Earth bow shock for many years now, observing its highly this area of . Essentially, MHD shocks nonstationary behaviour (see e.g. Lobzin et al. 2007). Another have been described using one of the following approaches, each prominent set of shock-related data was taken by the Voyager 1 of which possesses its own, inherent flaws. spacecraft, which, in late 2004, crossed the solar wind termina- First, from the theoretical MHD side, there are tion shock (Stone et al. 2005; Decker et al. 2005): the spacecraft Rankine-Hugoniot-like jump conditions, based on the con- observed power-law spectra of practically unchanging power in- servation of moments such as the MHD mass and dices across the shock (Cummings et al. 2006), in addition to a fluxes, the energy flux, and the conservation of the Poynting series of magnetic bumps and holes in the heliopause (Burlaga vector flux (see, e.g. Hudson 1970; Erkaev et al. 2000; Vogl et al. 2006a,b), which could be explained in terms of plasma et al. 2003). This approach is based on MHD and describes only waves. Up to now, however, these experimental data can only be a few low-order velocity moments of the plasma flow instead of fitted using by an ad hoc model because a robust physical expla- the full ion velocity distribution function f (w). Therefore, this nation of their origin does not exist yet. approach is well suited to the analysis of Maxwell-Boltzmann- A third approach to the understanding of understand MHD like distribution functions, where the entire function can be shocks has also emerged, that is based on numerical shock sim- parameterised using only a few velocity moments, but may be ulation models completed using powerful supercomputers (e.g. inappropriate for, say, power-law distribution functions, as those Hada et al. 2003; Scholer et al. 2003): this, in principle, allows found in most cosmic ray spectra (see, e.g. Schlickeiser 2002; to study not only the behaviour of the individual particles, but Fisk & Gloeckler 2006). In addition, it turns out that, for an also the nonstationarity in the “fine structure” of the transition anisotropic plasma, the jump conditions are underdetermined, region (i.e. fields and distribution functions inside the shock). where one possible parameterisation is to assume that the down- However, numerical simulations always require the introduction stream pressure anisotropy, λ2 = p⊥,2/p,2, is the free parameter of boundary conditions, in addition to a numerically stable “step- (Erkaev et al. 2000). Since an anisotropic plasma may emerge ping scheme”, which, in principle, may introduce additional, in the presence of magnetic fields (Chew et al. 1956), this unphysical terms in the underlying equations (see, e.g. Press restriction can be traced back to the inherent MHD flaw, where 1987–2002, Chap. 19). A consistent description of the physi- one initially obtains an infinite hierarchy of moment equations cal boundary conditions typically requires the simulation of a

Article published by EDP Sciences 328 M. Siewert and H.-J. Fahr: A Boltzmann-kinetical description of an MHD shockwith arbitrary field inclination system that is much larger than the transition region of the shock as charge exchange, ionization or wave-particle induced fric- itself, and resolving this region can be difficult. One example tion. Following the earlier studies of Fahr & Siewert (2006) of such a situation is the heliospheric termination shock, where and Siewert & Fahr (2007a,b), a sufficiently general form of the the heliosphere with an average radius of the order of 100 AU Boltzmann-Vlasov equation is given by requires a grid size that is considerably different from the typ- ical size of a shock, which is assumed to be of the order of d d As(w, s) f (w, s) = A(w, s) f (w, s) a few gyroradii (Scholer et al. 2003; Fahr & Siewert 2007). ds dw For these reasons, in any large-scale simulations, it is possible d to verify only the existence of a shock by using the Rankine- + A⊥(w, s) f (w, s), (1) dw⊥ Hugoniot jump conditions. Results from such numerical simula- tions demonstrate a highly nonstationary behaviour, similar to in where the basic has been specialised to a situ space observations, for which, however, no complete theo- spatially one-dimensional problem with the streamline coordi- retical understanding exists. nate s.Thecoefficients A and A⊥ are, at this point, free func- Since all of these approaches are, in one way or another, tions, which need to be derived from the physical processes in incomplete, we developed an approach based on a kinetic the transition region of the shock. Boltzmann-Vlasov equation, which might be able to fill the In a more general approach, one has to consider small fluctu- gaps between the competing descriptions. In contrast to common ations, in terms of plasma waves and turbulence. In the presence MHD, our approach provides a description of the entire distribu- of a stationary background magnetic field, these processes are tion function f (w), which includes all forms of nonthermal dis- described traditionally using the Fokker-Planck equation (see, tribution functions. So far, this model allowed us to explain the e.g. Schlickeiser 1989; Chalov & Fahr 1998),whichthenintro- conserved power-law index observed by the Voyager 1 space- duces terms such as craft (Siewert & Fahr 2007a,b), as well as the magnetic bumps and holes observed by the same spacecraft in the heliosheath 1 d 2 d w Dww f · (2) (Burlaga et al. 2006b; Fahr & Siewert 2007). Our results imply w2 dw dw that, to obtain a (quasi-)stationary transition region, one requires additional physical processes in addition to the deceleration of Obviously, these terms lead to the emergence of second-order and the change in electric and magnetic fields across the derivatives in Eq. (1), shock. Since, in our approach, the conservation of the mass flow d d transforms into a set of specific mathematical conditions, the ad- As(w, s) f (w, s) = A(w, s) f (w, s) ditional physical processes to be included must be of a rather ds dw specific form as well, unless the transition region is highly non- d + A⊥(w, s) f (w, s) stationary (Siewert & Fahr 2007b). This result is agrees with the dw⊥ nonstationarity emerging in both numerical approaches and ex- d2 + A  w, s f w, s perimental observations, which also implies that additional, in- , ( ) 2 ( ) ternal microphysics should be taken into account. In this study, dw we restrict ourselves to the solar wind termination shock, which 2 + w d w is located approximately at a similar spatial position over a sig- A⊥,⊥( , s) f ( , s) dw2 nificant period of time. We reinvestigate our kinetic model of ⊥ the MHD shock, and derive a new, improved form of the kinetic d2 + A,⊥(w, s) f (w, s), (3) Boltzmann equation for the general shock, which is based on dw dw⊥ a new, systematic connection between the MHD view and the per-particle view of the shock. These arguments enable to be as well as to further contributions to A and A⊥. For this rea- ffi removed most of the physical and mathematical problems that son, we call the coe cients Ai the first-order and second-order emerged in previous studies (e.g. Fahr & Siewert 2006; Siewert Fokker-Planck terms (which are not to be confused with the bet- & Fahr 2007a,b). As a side-result, we prove the equivalence of ter known Fokker-Planck coefficients). However, in this study, MHD invariants and single particle invariants, by generalising we focus on the more basic situation of a wave-free plasma, the conservation of the “magnetic CGL-moments”. where all second-order terms vanish, since in this special sit- uation, the general form of the solution is already known (see Siewert & Fahr 2007b, with the main mathematical methods re- 2. The kinetic approach peated later in this paper). Finally, we work with the gyroaver- aged approach commonly found in literature, i.e. there are only 2.1. The mathematical and physical decomposition two velocity components, the parallel velocity w,andtheper- of the Boltzmann equation pendicular velocity w⊥. Depending on the specific structure in- In Siewert & Fahr (2007b), we proved that the specific form side the transition region, this approach may be no longer justi- of the Boltzmann equation which was used in earlier studies fied; however, a more specific description of this configuration is inherently unable to describe the parallel MHD shock in the does not yet exist in an analytical approximation. anisotropic CGL model. For this reason, we now rederive the The Boltzmann equation may also be parameterised from a Boltzmann equation, looking carefully for shortcomings and physical standpoint that emphasizes not the mathematical form flaws in the up-to-now description, including missing terms. but the physics behind it. We write this alternate representation We consider a more general form of the Boltzmann equa- in the form tion that we later develop to include nontrivial terms, such as L [ f ] + L [ f ] wave-turbulence generation, diffusion and an exchange of par- kin F + ticle number, energy and momentum between multiple parti- Lacc[ f ] ∗ cle populations, which occur when we consider processes such + Qsource[ f, f ] = 0. (4) M. Siewert and H.-J. Fahr: A Boltzmann-kinetical description of an MHD shockwith arbitrary field inclination 329

Here Li is a linear differential operator acting on the distribution inside the shock is not arbitrary, but restricted by several pseudo- function; the subscripts stand for kinetic terms on the individual MHD properties. particle level, “true” Force terms, and pseudo forces within an In principle, there are two different ways to derive the factors accelerated reference frame. Finally, we take into account of a dwi/dt in their final form. There is first the per-particle approach, possible source/sink term, which represents a possible exchange where no a priori limits or averages are made, that considers a of particles, energy, and momentum between different particle single particle, with three velocity coordinates including a gy- populations; this term may be, in principle, nonlinear and de- roangle, and the full force acting on this particle. Then, one has pendent on multiple distribution functions, such as ions and to derive the modifications to this force caused by the local field . We denote the distribution function related to this ad- modifications inside the shock, which in turn have to be derived ditional population with an asterisk. In principle, the correspond- from the particle behaviour inside the system. In a sufficiently ing operator does not need to be linear, and is referred as Q. general case, this system of equations is complicated and has to be solved numerically, i.e. the analytical forms of the kinetic terms cannot be derived. 2.2. The Fokker-Planck terms in the various reference frames The other approach is what we call the semikinetic approach. From the kinetic view, we borrow the idea of taking an individ- A general formulation of the collisionless Boltzmann equation ual particle (or, alternatively, a narrow region in velocity space), (i.e. the Vlasov equation) in an accelerated reference frame is but we describe the force terms using MHD quantities. In other given by words, force terms and their corresponding energy and momen- tum exchange are implicitly included in MHD quantities such dU as the bulk velocity, the partial pressures, and the magnetic field (w · ∇x) f + (F · ∇w) f + · ∇w f = 0, (5) dt tension. These quantities must then be parameterised in a way that is consistent with the MHD jump conditions, which leads where dU/dt is the acceleration of the bulk plasma system that to a consistent description of the transition region using MHD moves with the bulk velocity U, i.e. this term may be identi- quantities only, where only a small part of the MHD parameters fied with L . The first term of the left-hand side of the ex- acc actually needs to be modelled. This produces, however, a com- pression corresponds to the kinetic term L , while the middle kin plicated equation because all important MHD quantities must be term may be identified with the electromagnetic force term L . F,em represented, that is the magnetic field, partial pressures, bulk ve- Specializing to a one-dimensional problem, we obtain locities, and mass density, which are all related in some way. d Finally, one requires a systematic relation between the MHD L [ f ](w, s) = w cos Θ f (w,w⊥)(6) kin Bn ds quantities and the individual kinetic velocities, unless all par- ticles react in an identical way to the shock, which would be Θ in the (accelerated) plasma frame, where Bn is the angle be- unphysical. In the remaining part of this section, we introduce tween the magnetic field and the shock normal. Since this is the a formalism than enables this to be realised in a straightforward only term containing spatial gradients of the distribution func- way. tion, one obtains easily the final form of As, We begin with a phenomenological motivation of this for- malism. First, we would like to emphasize that, in a wide variety As = w cos ΘBn(s) = Un(w, s), (7) of physical systems, the magnetic moment of the individual par- where we introduce the differential bulk velocity Un(w, s). This ticles is conserved, i.e. means that any temporal derivative appearing in the equation 2 may be transformed into a spatial derivative using d d w⊥ μ ∝ = 0. (10) d d dt dt B = w cos ΘBn(s) , (8) dt ds For a broad distribution function f (w), the total magnetic mo- Then, as long as the Fokker-Planck terms A and A⊥ are pro- ment of all particles is conserved, which is given by portional to d/dt, is it possible to completely eliminate the factor 2 w cos Θ (s). We note that this point was missing in an earlier d d w⊥ Bn μ = d3w f (w) = 0. (11) equation derived in Fahr & Siewert (2006), which produced a dt f dt B singularity under specific conditions. Deriving the “real” force term is more complicated because the electromagnetic tension However, this expression may be interpreted in terms of an MHD forces (i.e. electric and magnetic force terms) must be consid- velocity moment, i.e. ered. One possible alternative approach to this problem is d = d 1 3 2 w d f dw ∂ f dw⊥ ∂ f μ f d ww⊥ f ( ) = + , (9) dt dt B dt dt ∂w dt ∂w⊥ i i i i d 1 2 = w⊥ where the sums are performed over all parameters which get dt B d p⊥ modified by the shock (and no explicitly time-dependent terms = · (12) do appear). Unfortunately, collecting these parameters is still dt B rather complicated, as several nontrivial points have to be taken As a generalisation of this, any individual particle invariant that into account. For example, in pure MHD, the electromagnetic i j fields are considered to be frozen into the plasma on the far up- is proportional to w w⊥ may be trivially transformed into an stream and downstream sides of the plasma, which implies that MHD form, simply by applying this should be true also for the small transition region in which i j i j pure MHD no longer holds. Thus, the behaviour of the particles w w⊥ →w w⊥. (13) 330 M. Siewert and H.-J. Fahr: A Boltzmann-kinetical description of an MHD shockwith arbitrary field inclination

For the magnetic moment, the corresponding MHD invariant is In Appendix A, we demonstrate in detail how Eq. (22) is valid given by under the conditions presented at the start of this section. Although we do not study the physical nature of Eq. (21), we d p⊥ = const. (14) emphasize that the mostly mathematical approach to this iden- dt B tification requires that such an invariant must exist. In Fahr & Now, it is known that, in the CGL model for anisotropic MHD Siewert (2008), we identified this invariant in the solar wind, fol- systems (Chew et al. 1956), one obtains two adiabatic invariants lowing from the divergence of the plasma stream and the corre- that relate the different MHD parameters with each other. These sponding Parker model for the frozen-in magnetic fields (Parker two invariants are given by 1965). Using a general expression for the corresponding velocity modification, d d μ f d p⊥ CGL1 = = = 0, (15) d B · ∇ U · B dt dt ρ dt Bρ w = −w , (23) dt B B and the reason for the resulting change of the parallel velocity com- 2 d d pB CGL = = 0. (16) ponent w is due to the recognition of the bulk velocity gradient dt 2 dt ρ3 parallel to the magnetic field B, by the particle, when it covers apathw per unit time of its motion parallel to B.However, Obviously, the first CGL invariant is given by a “normalised” evaluating this expression requires an intimate knowledge of the form of the MHD magnetic moment. reaction of the frozen-in magnetic field, for which, inside the In an inertial rest frame, the normalisation of the station- = transition layer of the shock, no systematic theory yet exists. ary distribution function is constant, i.e.ρ ˙ 0. In our formal- Therefore, lacking any better description, we make the ad- ism, however, we are working in an accelerated reference frame,  hoc assumption that both CGL invariants are conserved inside whereρ ˙ 0. Introducing the normalised distribution function the shock. This approach requires that the magnetic field must f (w) be changing slowly. In other words, the reorientation and con- fnorm(w) = , (17) densation timescale τ must be much larger than the gyration ρ c timescale, we see that the perpendicular pressure is given by dB τc = /B τgyr. (24) 2 3 2 dt p⊥ ∝w⊥∝ d ww⊥ fnorm(w), (18) Since all MHD quantities in the system are connected with each and the additional factor ρ−1 in Eq. (15) cancels out the con- other, similar conditions must hold for the other MHD quantities tributions from the acceleration of the system. In other words, appearing in the adiabatic invariants as well. From this require- Eq. (13) has to be generalised to ment, it is automatically possible to derive another condition, namely the absence of particle-particle collisions. Since the con- i j i j servation of the magnetic moment requires slow variations of all w w⊥ w w⊥   i j → = w w⊥norm, (19) parameters, including the perpendicular particle velocities, any ρ ρ fast energy and momentum exchange mechanism (such as scat- which is valid both in inertial and non-inertial rest frames. tering) must be absent from the system, or the gyration of the Next, we demonstrate that, in the absence of stochastic pro- individual particles would be significantly perturbed. The same cesses, the opposite direction of Eq. (19) is also valid. In other requirement holds for the second CGL invariant to be valid (see words, it is possible, under certain conditions, to interpet an e.g. Kulsrud 1983, p. 115). MHD invariant in terms of an integral over per-particle invari- Finally, we consider the Eqs. (21)and(10)toevaluatethe ants. Taking, representatively, the second CGL invariant, we may temporal derivative and obtain the expressions write dw d = w (ln ρ − ln B) (25) d p B2 d B2 dt dt = d3ww2 f (w) CGL2 dt ρ ρ2 dt  ρ2 norm and d B2 = d3w f (w) w2 dw⊥ w⊥ d norm dt  ρ2 = ln B. (26) dt μ 2 dt 2 B 2 d 3 + w d w fnorm(w) = 0, (20) These two Eqs. (26)and(25), are sufficient to describe the sys- ρ2 dt tem. Then, collecting our results, the Fokker-Planck terms are Obviously, as soon as the second integral in this sum vanishes, given by the first integral vanishes as well, which requires that = Θ As w cos Bn (27) 2 d 2 B 2 d w = 0. (21) A = −w cos Θ (ln ρ − ln B) (28) dt ρ2  Bn ds w⊥ d Since this approach is not restricted to the second CGL invariant A⊥ = − w cos ΘBn ln B. (29) alone, we may write our condition in the more general form 2 ds In these equations, the additional negative sign follows from the  i j  i j d ww⊥ ! d ww⊥ fact that, in Eq. (1), the force term was moved to the other side = d3w f (w) = 0. (22) dt ρ norm dt ρ of the equation. From the argument following Eq. (8), the global M. Siewert and H.-J. Fahr: A Boltzmann-kinetical description of an MHD shockwith arbitrary field inclination 331 factor w cos ΘBn may be removed, which allows us to write the i.e. the statistical weights f (w) at the individual points in velocity simpler form space are simply moved around, but not smeared out. Here, the subscripts 1 and 2 denote the upstream and downstream distribu- = As 1 (30) tion functions, respectively. Next, we take into account that the d absolute normalisation of the distribution function in an acceler- A = −w (ln ρ − ln B) (31) ds ated rest frame is not constant, which is conventionally parame- = w⊥ d terised using the MHD compression ratio x,i.e.ρ2 xρ1. Then, A⊥ = − ln B. (32) we write down this relation using the full distribution functions, 2 ds Then, we obtain the final form of the kinetic Boltzmann equa- 3 ρ2 = d w¯ Df1(¯w, w¯ ⊥) tion, ! 3 d d d w⊥ d d = xρ1 = x d w f1(w,w⊥), (37) f = −w (ln ρ − ln B) f − ln B f. (33) ds ds dw 2 ds dw⊥ where x is an arbitrary positive number that has to be derived This equation describes the full downstream distribution func- from the MHD jump conditions. Since this relation must not de- tion at an MHD shock (instead of only a few, low-order veloc- pend on the fi or x, we see that ity moments) using the assumption that the adiabatic invariants of the CGL theory also hold inside the transition region of the d3w¯ D =! d3w x. (38) shock. It it worth mentioning that this equation does not depend In other words, the Jacobi determinant of the transformation upon the magnetic field orientation, which is represented by must encode the compression ratio between the upstream and the fact that no magnetic field projections (Bn or Bt) appears. downstream sides in a specific way. This result means that D Therefore, the terms related to this effect derived by Fahr & may not depend upon the particle velocities, and the relation Siewert (2006) must be discarded on account of mixing the connecting upstream and downstream variables must therefore semikinetic approach with the full kinetic approach. Since both be linear, such as approaches rely on different amounts of averaging and other ap- = + proximations, a self-consistent description of the shock must not w Cw¯  Cw¯ ⊥ (39) mix these different representations. w⊥ = C⊥w¯  + C⊥w¯ ⊥. (40) Now, all that remains to do is to determine the form of the coeffi- 2.3. Solutions of the improved Boltzmann equation cients Ci and Bi.Todothis,wetakeEq.(35) on the downstream side and express it in the integral form, In Siewert & Fahr (2007b), we derived restrictions for the possi- ble form of the Fokker-Planck terms Ai, based on the concept ! 3 that the average parallel velocity w vanishes in the plasma 0 = d ww f2(w,w⊥) frame. At this point, we repeat the nontrivial parts relevant for 3 the solution of our kinetic equation. First, we emphasize that the = d ww f1(¯w(w), w¯ ⊥(w)) adiabatic invariants (Eqs. (15)and(16)) are only valid in the rest frame comoving with the system. This may be understood 3 = d w¯ D (Cw¯  + C w¯ ⊥) f (¯w, w¯ ⊥) since we derived the Fokker-Planck terms A and A⊥ using adi-  1 abatic invariants depending on the partial pressures p and p⊥. 3 Conventionally, these partial pressures are taken in the “natural” = 0 + B d w¯ Df1(¯w, w¯ ⊥). (41) rest frame that is comoving with the plasma, since the integral Here, the first term vanishes because of our choice of the ref- w2 = dww2 f (w) (34) erence frame, while the second integral is always greater than zero. Therefore, to make the entire expression vanish, one re- quires C = 0, i.e. the parallel velocities do not become mixed is not invariant under a coordinate transformation of the form  → + with the perpendicular ones. Using Eq. (38) and writing out the w w U. Now, this specific reference frame is characterised ff by the fact that the velocity moment w vanishes. Taking our di erential then allows to prove that the perpendicular veloci- specific choice of the velocity coordinate system, this translates ties do not get parallel contributions either. This means that the into the requirements that most general, mass flow conserving transformation between the upstream and downstream coordinates is given by w ! w = Cw¯  (42) w = w⊥ cos φ = 0, (35) = w⊥ sin φ w⊥ C⊥w¯ ⊥. (43) where the second and third terms vanish because the distribu- Then, using Eq. (38), one immediately sees that tion function f does not depend upon the gyroangle φ. The only 3 3 2 3 d w¯ D = d w CC⊥ = d w x, (44) nontrivial part of this equation is w = 0. 3 In the absence of stochastic processes, any single point in where the additional factor C⊥ follows from the fact that d w ∝ phase space will remain forever a single point, which justifies w⊥ in cylinder coordinates. In other words, we obtain the addi- the approach tional condition

2 f2(w,w⊥) = f1(¯w(w), w¯ ⊥(w)), (36) CC⊥ = x. (45) 332 M. Siewert and H.-J. Fahr: A Boltzmann-kinetical description of an MHD shockwith arbitrary field inclination

Now, taking Eqs. (36) and the most general coordinate transfor- Table 1. Initial upstream and final downstream parameters for a single- mation (i.e. Eqs. (42)and(43)), we see that the most general, fluid, ion-only plasma. The downstream pressure anisotropies λ are not mass flow conserving solution of Eq. (1)isgivenby estimated, but are an exact result using Eq. (54). Normalised values are used. w ⊥ = w f2(w,w⊥) f1 , , (46) Upstream Downstream C C⊥ Perpendicular shock where the parameters Ci must be independent of w. ρ, x 1 2.0078 5 × 4 From this point, the rest of our formalism is rather straight- Un 10 4.9805 10 forward mathematics. Taking Eqs. (42)and(43), inserting them λ 1 2.0078 into Eq. (1) and comparing coefficients then allows to reduce p 0.01 0.066 p⊥ 0.01 0.133 the partial differential equation to the two ordinary differential ◦ Inclined (ΘBn = 45 ) equations ρ, x 1 2.137 5 × 4 dC A Un 10 4.679 10 = − C (47) λ 1 1.718 ds w p 0.01 0.107 p⊥ 0.01 0.183 and ◦ Almost parallel shock (ΘBn = 1 ) ρ, x 1 2.253 dC⊥ = − A⊥ C⊥. (48) U 105 4.4370 × 104 ds w⊥ n λ 1 0.1971 Since the coefficients Ci must not depend upon w, the Fokker- p 0.01 0.567 Planck terms Ai must be linear functions of wi, p⊥ 0.01 0.112

Ai(wi, s) = ai(s)wi, (49) ff and the di erential equations are formally solved by while for the perpendicular shock (ΘBn = π/2), where one ob- tains B2 = xB1 (Erkaev et al. 2000), the downstream pressure Ci = exp − ai ds . (50) anisotropy is given by = Then, the Eqs. (49), (50)and(45), applied in this order, allow λ2,⊥ xλ1. (56) us to determine if a kinetic equation is able to describe an MHD shock while conserving the mass flow, and the downstream dis- Obviously, the kinetic behaviour of an arbitrary distribution tribution function is given by Eq. (46). function across the shock depends on only two parameters, the Now, we may apply this formalism to the equation that MHD compression ratio x and the total magnetic field strength we derived in this study. Obviously, the Ai given by Eqs. (31) ratio B2/B1, i.e. on one parameter related to the massive parti- and (32) fulfill Eq. (49), i.e. they are linear functions of their cles, and one parameter related to the electromagnetic fields. It corresponding velocities. Since they are of the form d/ds ln g(s), is noteworthy that the behaviour of the distribution function de- evaluating Eq. (50) is trivial as well because the exponential and pends only on upstream and downstream quantities, and not on logarithmic functions cancel out, leading to the fine structure of the shock itself. In conventional MHD, it is usually assumed that a few velocity moments of low order are B1 ffi C = x (51) su cient to describe the behaviour of the system; in this light, B2 our result may be interpreted in a way that, at least when con- sidering shocks, these few lowest-order moments may be further and boiled down to a single kinetic parameter. This result likewise B2 hints that our kinetic equation successfully describes the non- C⊥ = , (52) B MHD region of the shock in a quasi-MHD approximation, which 1 is as close to MHD as possible, while leaving a sufficient number from which we automatically see that Eq. (45) is indeed fulfilled. of degrees of freedom to include strictly non-MHD behaviour. We may also derive analytic relations between all upstream and We return to this point in a future study. downstream MHD moments in the plasma rest frame, which are Since this formalism was derived in the comoving reference equivalent to knowledge of the distribution function. According frame of the plasma, which must always exist no matter how to Siewert & Fahr (2007b), this relation is given by complicated the microphysics in the system may be, we call this i j 2 i j i j the minimal kinetic extension. In other reference frames, this aij,2 = ww⊥ = (CC⊥)CC⊥aij,1 = xCC⊥aij,1. (53) 2 extension should be similarly applicable, although the transfor- mation of the velocity moments between the different reference The downstream pressure anisotropy derived using this formal- frames is mathematically complicated, and not pursued further ism is then in this study. C2 3 Using the treatment of the MHD shock presented in this pa- = ⊥ = B2 −2 λ2 λ1 x λ1. (54) per, we show a few selected results. We represent the upstream C2 B  1 distribution function using a bi-Maxwellian function,

For the parallel shock (ΘBn = 0), this relation simplifies to ⎛ ⎞ ⎜ w2 2 ⎟ ⎜  w⊥ ⎟ λ1 f (w) = exp ⎝⎜− − ⎠⎟ , (57) λ  = , (55) 2 2 2, x2 θ θ⊥ M. Siewert and H.-J. Fahr: A Boltzmann-kinetical description of an MHD shockwith arbitrary field inclination 333

1 with inclined, us 0.9 inclined, ds 2kTi 2pi 0.8 θi = = , (58) mp ρ 0.7 0.6 where the second identity follows from the ideal gas equation. Then, we use the same upstream parameters as those used by f 0.5 Erkaev et al. (2000), i.e. the Alfvenic Mach number MA = 2 0.4 and the sonic parameter As = 0.01. Using these parameters, it 0.3 is possible to derive a more conventional form of the upstream 0.2 parameters, the magnetic field strength 0.1 2 0 4πρ1U = 1 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 B (59) w Ma || = and the upstream perpendicular pressure, Fig. 1. Representative cuts through the distribution function at w⊥ 0 for the inclined shock, using an upstream bi-Maxwellian distribution = 2 function and the parameters from Table 1. p⊥ Asρ1U1. (60) In addition to the initial two dimensionless parameters, we use 1 inclined, us a more conventional approach to the upstream mass flow, by 0.9 inclined, ds −3 adopting the mass density ρ1 = 1m and the upstream bulk ve- 5 0.8 locity U = Un = 10 m/s. Since the solution of the anisotropic 0.7 jump conditions does not depend on the mass flow Unρ (see, e.g. Erkaev et al. 2000; Vogl et al. 2003), any other choice of the pa- 0.6 rameters constituting the mass flow would lead to similar results. f 0.5 In contrast to this, we remark that the dimensionless parame- 0.4 ters used earlier would influence the result. Finally, we assume that, on the upstream side, the plasma is perfectly isotropic, i.e. 0.3 λ1 = 1, and that the magnetic field orientation with respect to 0.2 the shock normal is characterised by the angle ΘBn. For conve- 0.1 nience, these upstream parameters are collected in Table 1. 0 Next, we solve the anisotropic MHD jump conditions 0 0.1 0.2 0.3 0.4 0.5 0.6 (Hudson 1970; Erkaev et al. 2000) for the upstream parame- w⊥ ters given. In a classical (i.e. MHD-only) approach to this prob- Fig. 2. Representative cuts through the distribution function at w = 0 lem, the equations are underdetermined, and one is faced with for the inclined shock, using an upstream bi-Maxwellian distribution one more downstream parameter than equations, which means function and the parameters from Table 1. that an additional equation must be derived using a different for- malism. In this study, we follow Erkaev et al. (2000)andse- lect the downstream pressure anisotropy λ2, which may be de- scribed using Eq. (54). Using this approach, we finally arrive at C⊥ = 1, which leads to a completely unmodified perpendicular a unique solution for the downstream parameters, which is given velocity. This effect is demonstrated in Figs. 3 and 4,wherewe in Table 1. Using this solution, we derive the MHD compression present, again, cuts through the upstream and downstream distri- ratio x and the “field compression ratio” B2/B1, allowing us to bution functions for an (almost) parallel shock. This is similar to determine the full distribution function on the downsteam side. the earlier result obtained for the perpendicular shock (Siewert Since the perpendicular shock has already been treated in & Fahr 2007a), where the parallel velocity components remain Siewert & Fahr (2007a), and the current approach to the force untouched. terms leads to identical results, we focus on the inclined and Finally, we point out that Eq. (33) does not depend upon parallel shocks. In Figs. 1 and 2, we present cuts through the the behaviour of the MHD quantities inside the transition re- ◦ distribution functions for an inclined shock (ΘBn = 45 )onthe gion, which is in excellent agreement with MHD, and which upstream and downstream sides, at w⊥ = 0andw = 0, respec- also implies that we indeed find a kinetic description for the tively. These figures demonstrate a basic property of our solu- MHD shock that depends essentially only on MHD quantities. tion (Eq. (46)), namely that the shock does not modify the basic However, we emphasize that this approach is, by no means, a shape of the distribution function, which is still of the character- globally complete description, but applies only under the re- istic Maxwellian form. Instead, it modifies the broadness of this strictions imposed by the MHD approach to the shock. First distribution, i.e. the parameter θi in Eq. (57), which is a func- of all, MHD requires that the electromagnetic fields are frozen- tion of the partial pressures. For the inclined shock, both compo- in, i.e. convected along with the background plasma. Without nents of the velocity are modified, which directly follows from this requirement, there would be no motion perpendicular to the the fact that, in this case, both coefficients Ci are not unity (i.e. magnetic field lines, and no perpendicular shock either. in Eqs. (51)and(52), the magnetic field ratio is not 1 or x). Therefore, although the frozen-in field condition is derived For the parallel shock, the situation is different. Since this ap- within the framework of classical MHD, it must be valid even proach is defined by B = Bn, and it follows from the MHD jump when all other MHD requirements fail, since otherwise, there conditions that the normal magnetic field is conserved, the mag- would be no perpendicular shock. For this reason, it must be ex- netc field ratios appearing in the coefficients Ci are unity, and pected that the fields are still frozen-in into the system even in 334 M. Siewert and H.-J. Fahr: A Boltzmann-kinetical description of an MHD shockwith arbitrary field inclination

1 almost parallel, us describe wave generation, by including the wavemodes as yet 0.9 almost parallel, ds another separate fluid. On the kinetic level, such interactions 0.8 between various components of the model is realised by up- 0.7 grading the Boltzmann-Vlasov equation to a Fokker-Planck-like form, where the interactions are parameterised as diffusion coef- 0.6 ficients. To the best of our knowledge, there exists no compara- f 0.5 ble systematic theory of interacting fluids on the MHD level yet. 0.4 The closest thing to such a theory found in literature is two-fluid 0.3 hydrodynamics (see Holzer & Axford 1970). Clearly, a self-consistent solution to all of these problems 0.2 may become complicated and is far beyond the scope of this 0.1 present study. As a first step towards such a description, we are 0 currently working on a multifluid generalisation of the classical -0.6 -0.4 -0.2 0 0.2 0.4 0.6 MHD jump conditions. Although this work is close to comple- w|| tion, we point out that the “initial problem”, i.e. the fact that

Fig. 3. Representative cuts through the distribution function at w⊥ = the anisotropic jump conditions are not perfectly closed, appears ◦ 0 for the almost parallel shock (ΘBn 1 ), using an upstream bi- to be only the literal tip of the iceberg; for multiple fluids, the Maxwellian distribution function and the parameters from Table 1. amount of free parameters seems to be growing, which offers an excellent interface to include fluid-fluid interactions, in terms 1 almost parallel, us of additional conditions required to close the generalised jump 0.9 almost parallel, ds conditions. In face of all these aspects, our current result must 0.8 be interpreted as a working, self-consistent description of the classical, single-fluid MHD shock only, and as a basis for future 0.7 work. 0.6

f 0.5 3. Applications and outlook 0.4 0.3 3.1. On the incompleteness of a single-fluid system 0.2 Taking Eq. (53), we see that the partial downstream pressures 0.1 are given by 2 0 p = xC p (61) 0 0.1 0.2 0.3 0.4 0.5 0.6 ,2  ,1 w⊥ and Fig. 4. Representative cuts through the distribution function at w = 2 ◦ p⊥,2 = xC⊥ p⊥,1. (62) 0 for the almost parallel shock (ΘBn 1 ), using an upstream bi- Maxwellian distribution function and the parameters from Table 1. As demonstrated by Erkaev et al. (2000), the anisotropic MHD jump conditions are underdetermined, and one additional equa- tion is required, which they associated with the downstream the transition layer of the shock, in the sense of a generalised pressure anisotropy λ2. Now, however, the minimal kinetic ex- frozen-in field condition. tension gives us two additional equations, transforming the In addition, the MHD approach to shocks requires that the underdetermined system of equations into an overdetermined system is charge-neutral, i.e. that there are no local electric cur- one. This follows from the fact that the downstream pressure rents present. However, in a more consistent description, one has anisotropy is defined by to include both ions and electrons as separate, interacting fluids. On the other hand, the jump conditions are explicitly tailored p⊥,2 λ2 = , (63) to one single fluid, which is typically interpreted in terms of an p,2 ion flow, with the implicit asumption that the much lighter elec- trons are convected along with the rest of the system, and that which is invariant under the transformation quasineutrality is obtained. Therefore, to arrive at a more consis- p ⊥ = C · p,⊥, (64) tent description of the MHD shock, one requires a two-fluid gen- , eralization of the MHD jump conditions, including an MHD for- indicating that the absolute normalisation of the pressures is not mulation of charge-neutrality on the upstream and downstream preserved by λ2,andthat,takingλ2 alone might result in a so- sides. This, however, opens yet another problem, namely the lution that does not satisfy Eqs. (61)and(62). Introducing the fact that, inside the transition region, where MHD is not ap- parameter (Erkaev et al. 2000) plicable, quasineutrality may no longer be an absolute require- p − p⊥ = − π , ment. Considering that the distribution function may 1 4 2 (65) be quite different from the ion one, this is clearly not a trivial B problem. We emphasize that even the particle-field interactions one may express most of the partial pressure terms in the jump present in MHD may already be interpreted as a two-fluid sys- conditions as a function of and λ2, with one isolated perpen- tem, with one fluid being composed of massive particles, and dicular pressure remaining, which enables us to derive one of the other fluid of frozen-in fields. In light of this interpretation, the partial pressures from MHD, and leads to two determina- interactions between multiple fluids should, in fact, be possible tion conditions for p⊥,2.However, is invariant under Eq. (64), in the framework of MHD. Such an approach would allow to since such a renormalisation may also be interpreted in terms of M. Siewert and H.-J. Fahr: A Boltzmann-kinetical description of an MHD shockwith arbitrary field inclination 335

B 2 → C · B2.Thisissufficient to prove that Erkaev et al. (2000) which is required to allow motion perpendicular to the mag- are unable to predict the correct normalisation of the partial pres- netic field, as it is allowed in conventional MHD. Clearly, the sures, and that just providing an expression for λ2 alone is insuf- time-dependent term, which is usually set to zero on the far ficient to arrive at a closed system of equations. This may be un- upstream and downstream sides, must be nonzero inside the derstood since any theory capable of predicting the downstream shock to preserve frozen-in fields. Since the presence of time- pressure anisotropy must also be able to predict both individual dependent terms in electrodynamical equations is usually inter- partial pressures, which results in two more equations instead preted in terms of plasma waves, this automatically hints that of just one, replacing the previously underdetermined system of an MHD shock is a natural plasma wave generator. However, equations with an overdetermined system. A possible solution since plasma waves usually require a quasineutral system (i.e. for this situation is the inclusion of electrons, which introduces where ions and electrons are present in equivalent quantities), additional equations and parameters that might lead to a more one needs to extend the MHD jump conditions to include at least consistent description. Based on our current description of a sin- two particle flows (or, alternatively, two charge flows). As al- gle fluid shock, we are currently working on a consistent de- ready mentioned, we are working on a consistent description of scription of a multifluid shock, which may be used to explicitly all these aspects. We expect that this description is able to de- model quasineutrality and possibly also stochastic interactions scribe the behaviour of the magnetic fields inside the transition or hybrid fluid-particle descriptions in a more systematic way region, shedding more light on the currently unsolved points re- than commonly found in literature. lated to the second CGL invariant.

3.2. The parallel shock and the transition region 4. Conclusions Taking the general anisotropic jump conditions (Erkaev et al. = = In this study, we derived an improved version of an earlier ki- 2000) and specialising them to the parallel shock (Bt 0, Bn netic Boltzmann equation derived by Fahr & Siewert (2006), const.), one obtains which attempts to describe MHD shocks, such as the solar wind termination shock. Using a more strict approach in terms of ref- [[B ]] = 0 (66) n erence frames and initial assumptions, we were able to eliminate = [ρUn] 0 (67) the restrictions which emerged in the earlier studies, strength- = [[UtBn]] 0 (68) ening the connection between MHD and kinetic theory. This 2 new equation fulfils the requirement derived by Siewert & Fahr [p + ρU ] = 0 (69) n (2007b) based on the conservation of the mass flow, which sug- [ρU U ] = 0 (70) n t gests that Eq. (33) is a self-consistent description of a basic, 2 turbulence-free MHD shock that depends only on MHD up- 3 Un Un p + p⊥ + = 0, (71) stream and downstream quantities, but not on the behaviour of 2 2 the plasma in the transition region. Obviously, this relation does not contain any variable magnetic In addition, we derived what might turn out to be a new the- field terms, which may be interpreted in terms of the plasma ory of per-particle invariants, derived from MHD invariants, gen- following the field lines. In other words, pure MHD is unable eralising the well-known equivalence between the single-particle to describe a parallel shock, and inside the transition region of magnetic moment conservation and the equivalent MHD adia- the shock, pure MHD must no longer hold. The existence of a batic invariant. While we have not yet been able to prove that parallel shock is mostly accepted on the condition that many as- this generalisation leads to physical expressions for all possible trophysical shock configurations (see e.g. Treumann & Scholer MHD invariants, we have found several arguments that strongly 2002) do require this configuration. From a mathematical point suggest that this approach works, at least, for the two adiabatic of view, the parallel shock may be described in terms of the invariants appearing in the CGL theory. Our current work hints ◦ that the conservation of the second CGL invariant is related to a limit ΘBn → 0 . As it turns out, the downstream transversal mag- netic field does not converge towards zero in this limit, and the bulk velocity gradient parallel to B and the corresponding reac- perfectly parallel shock may be unphysical, being replaced in- tion of the frozen-in magnetic field (Fahr & Siewert 2008). stead with a shock where the magnetic field is parallel on the upstream side, but not on the downstream side. However, this Acknowledgements. We are grateful for financial support to the DFG within the approach requires considerably more work, as the conventional frame of the DFG-Project Fa 97/31-2. anisotropic Rankine-Hugoniot equations do not allow such a so- lution. While, in principle, many interpretations of this behaviour Appendix A: Transformation of an MHD invariant are possible, perhaps the most straightforward idea is that the in a per-particle invariant transition region of the MHD shock differs from ideal MHD pre- dictions, and that some of the jump conditions have to be modi- In this appendix, we prove that Eq. (22) is always fulfilled, for fied. Since energy and momentum are conserved quantities even arbitrary distribution functions f1(w). Writing down the temporal outside of MHD, and the normal magnetic field conservation is derivative of this expression, one obtains the following require- related to many other, non-MHD plasmaphysical applications as ment well, the conservation of the transverse electric field is the only 3 d i j d i j ! MHD jump conditions which may, perhaps, be modified by the d w f1(w) w w⊥ + f1(w) w w⊥ = shock. This jump condition is closely related to the so-called dt dt frozen-in field condition (Alfvén & Fälthammar 1963), 3 d i j − ∇ × × = d w f1(w) w w⊥ . (A.1) ∂t B (U B) 0, (72) dt 336 M. Siewert and H.-J. Fahr: A Boltzmann-kinetical description of an MHD shockwith arbitrary field inclination

In other words, we require Here, the second term vanishes because of Eq. (A.6). The first term may be removed by noting that, when controlled by the 3 d i j ! collisionless Boltzmann-Vlasov equation, individual particles in d w f (w) w w⊥ = 0. (A.2) dt 1  a physical system do always follow determined trajectories, and that therefore the weight functions αn(t) are constant parameters To prove this, we begin by using the most simple distribution on these trajectories as a consequence of Liouvilles theorem. For function, this reason, the first term also vanishes, and we have proven that f (w) = δ(w − w ), (A.3) any temporal change of a velocity moment in an N-particle sys- 0 tem may be described using the sum of changes of the individual which describes a single particle (or a single “cell” in velocity particles. space). Inserting this into Eq. (A.2) leads to This approach works only in the absence of stochastical pro- cesses, which destroy the uniqueness of the particle trajectories. d d3w δ(w − w ) wi w j =! 0. (A.4) For this reason, it may be assumed that this approach works for a dt 0  ⊥ broad distribution function only when stochastical processes are still absent, as it is the case for the Boltzmann-Vlasov equation. Here, the derivative of the delta function trivially vanishes be- We will present a more detailed analysis under which conditions cause of this holds in a future publication. d d d3wδ(w − w ) wi w j = wi w j 0  ⊥ ,0 ⊥,0 References dt dt Alfvén, H., & Fälthammar, C. G. 1963, Cosmical Electrodynamics, 2nd Ed. 3 d i j = d wδ(w − w0) w w⊥. (A.5) (Oxford: Clarendon Press) dt Burlaga, L. F., Ness, N. F., & Acuna, M. H. 2006a, ApJ, 642, 584 Burlaga, L. F., Ness, N. F., & Acuna, M. H. 2006b, Geophys. Res. Lett., 33, This automatically means that L21106 Cercignani, C. 1988, The Boltzmann Equation and Its Applications (New York: Springer-Verlag) 3 d i j d w δ(w − w0) w w⊥ = 0. (A.6) Chalov, S. V., & Fahr, H. J. 1998, A&A, 335, 746 dt Chew, G. F., Goldberger, M. L., & Low, F. E. 1956, Proc. R. Soc. London A, 236, 112 For a system with N particles, the situation is more complicated. Cummings, A. C., Stone, E. C., McDonald, F. B., Heikkila, B. C., & Lal, N. Here, the distribution function is given by 2006, in Physics of the inner Heliosheath, AIP Conf. Proc., 858, 86 Decker, R. B., Krimigis, S. M., Roelof, E. C., et al. 2005, Science, 309, 2020 N Erkaev, N. V., Vogl, D. F., & Biernat, H. K. 2000, J. Plasma Physics, 64, 561 f (w) = α δ(w − w ), (A.7) Escoubet, C. P., Schmidt, R., & Goldstein, M. L. 1997, Space Sci. Rev., 79, 11 1 n 0,n Fahr, H.-J., & Siewert, M. 2006, A&A, 458, 13 n=1 Fahr, H.-J., & Siewert, M. 2007, ASTRA, 3, 21 Fahr, H.-J., & Siewert, M. 2008, A&A, 484, L1 where Fisk, L. A., & Gloeckler, G. 2006, ApJ, 640, L79 N Hada, T., Onishi, M., Lembege, B., & Savoinin, P. 2003, J. Geophys. Res., 108, = 1233 αn(t) 1(A.8)Holzer, T. E., & Axford, W. I. 1970, ARA&A, 8, 31 n=1 Hudson, P. D. 1970, Planet. Space Sci., 18, 1611 Kulsrud, R. M. 1983, in Handbook of Plasma Physics, ed. M. N. Rosenbluth, & and R. Z. Sagdeev (Amsterdam, North-Holland: Elsevier), 1, 115 N Lobzin, V. V., Krasnoselskikh, V. V., Josqued, J.-M., et al. 2007, d Geophys. Res. Lett., 34, L05107 α (t) = 0. (A.9) dt n Parker, E. N. 1965, Plan. Sp. Sc., 13, 9 n=1 Press, W. H. 1987–2002, Numerical recipes (New York: Cambridge University Press) Inserting this distribution function into Eq. (A.2) leads to Schlickeiser, R. 1989, ApJ, 336, 243 Schlickeiser, R. 2002, Cosmic Ray Astrophysics (Berlin: Springer Verlag) N Scholer, M., Shinohara, I., & Matsukiyo, S. 2003, J. Geophys. Res., 108, 1014 ! d 3 i j Siewert, M., & Fahr, H.-J. 2007a, A&A, 463, 799 0 = α (t) d wδ(w − w ) w w⊥ (A.10) n n  Siewert, M., & Fahr, H.-J. 2007b, A&A, 471, 7 = dt n 1 Stone, E. C., Cummings, A. C., McDonald, F. B., et al. 2005, Science, 309, 2017 Treumann, R. A., & Scholer, M. 2002, in The Century of Space Science d (Norwell: Kluwer Academic Publishers), 1495 +α (t) d3w δ(w − w ) wi w j . (A.11) n dt n  ⊥ Vogl, D. F., Langmayr, D., Erkaev, N. V., et al. 2003, Planet. Space Sci., 51, 715