1: %% Beginning of file 'sample.tex'
2: %%
3: %% Modified 03 Jan 01
4: %%
5: %% This is a sample manuscript marked up using the
6: %% AASTeX v5.x LaTeX 2e macros.
7:
8: %% The first piece of markup in an AASTeX v5.x document
9: %% is the \documentclass command. LaTeX will ignore
10: %% any data that comes before this command.
11:
12: %% The command below calls the preprint style
13: %% which will produce a one-column, single-spaced document.
14: %% Examples of commands for other substyles follow. Use
15: %% whichever is most appropriate for your purposes.
16:
17: \documentclass[12pt,preprint]{aastex}
18:
19: %% manuscript produces a one-column, double-spaced document:
20:
21: %%\documentclass[manuscript]{aastex}
22:
23: %% preprint2 produces a double-column, single-spaced document:
24:
25: %\documentclass[preprint2]{aastex}
26:
27: %% If you want to create your own macros, you can do so
28: %% using \newcommand. Your macros should appear before
29: %% the \begin{document} command.
30: %%
31: %% If you are submitting to a journal that translates manuscripts
32: %% into SGML, you need to follow certain guidelines when preparing
33: %% your macros. See the AASTeX v5.x Author Guide
34: %% for information.
35:
36: \newcommand{\vdag}{(v)^\dagger}
37: \newcommand{\myemail}{dastgeer@ucr.edu}
38:
39: %% You can insert a short comment on the title page using the command below.
40:
41: \slugcomment{ Preprint: {\it Astrophysical Journal}}
42:
43: %% If you wish, you may supply running head information, although
44: %% this information may be modified by the editorial offices.
45: %% The left head contains a list of authors,
46: %% usually a maximum of three (otherwise use et al.). The right
47: %% head is a modified title of up to roughly 44 characters. Running heads
48: %% will not print in the manuscript style.
49:
50: \shorttitle{Energy Cascades in Astrophysical Plasma}
51: \shortauthors{Dastgeer \& Zank}
52:
53: %% This is the end of the preamble. Indicate the beginning of the
54: %% paper itself with \begin{document}.
55: \newcommand{\itemfig}[1]{\item{\bf Fig (#1) }}
56: \newcommand{\be}{\begin{equation}}
57: \newcommand{\ee}{\end{equation}}
58: \newcommand{\et}{{\it{et. al. }}}
59: \newcommand{\eqs}[2]{Eqs. (\ref{#1}) \& (\ref{#2})}
60: \newcommand{\Eqs}[2]{Equations (\ref{#1}) \& (\ref{#2})}
61: \newcommand{\Eq}[1]{Equation (\ref{#1})}
62: \newcommand{\eq}[1]{Eq. (\ref{#1})}
63: \newcommand{\fig}[2]{Figs (\ref{#1}) \& (\ref{#2})}
64: \newcommand{\Fig}[1]{Fig. (\ref{#1})}
65: \newcommand{\eqa}{\begin{eqnarray}}
66: \newcommand{\eeq}{\end{eqnarray}}
67: \newcommand{\eqsto}[2]{Eqs. (\ref{#1}) to (\ref{#2})}
68: \newcommand{\Eqsto}[2]{Equations (\ref{#1}) to (\ref{#2})}
69: \newcommand{\sig}{\sigma_{\rm ex}}
70: \begin{document}
71:
72: %% LaTeX will automatically break titles if they run longer than
73: %% one line. However, you may use \\ to force a line break if
74: %% you desire.
75:
76: \title{Energy Cascades in a Partially Ionized Astrophysical Plasma}
77:
78:
79: %% Use \author, \affil, and the \and command to format
80: %% author and affiliation information.
81: %% Note that \email has replaced the old \authoremail command
82: %% from AASTeX v4.0. You can use \email to mark an email address
83: %% anywhere in the paper, not just in the front matter.
84: %% As in the title, you can use \\ to force line breaks.
85:
86: \author{Dastgeer Shaikh\footnote{Electronic mail : {\tt dastgeer@ucr.edu}}
87: \and
88: Gary P. Zank\footnote{Electronic mail : {\tt zank@ucr.edu }}}
89: \affil{Institute of Geophysics and Planetary Physics (IGPP),\\
90: University of California, Riverside, CA 92521. USA.}
91:
92:
93:
94:
95:
96:
97: %% Notice that each of these authors has alternate affiliations, which
98: %% are identified by the \altaffilmark after each name. Specify alternate
99: %% affiliation information with \altaffiltext, with one command per each
100: %% affiliation.
101:
102:
103:
104:
105: %% Mark off your abstract in the ``abstract'' environment. In the manuscript
106: %% style, abstract will output a Received/Accepted line after the
107: %% title and affiliation information. No date will appear since the author
108: %% does not have this information. The dates will be filled in by the
109: %% editorial office after submission.
110:
111: \begin{abstract}
112: A local turbulence model is developed to study energy cascades in the
113: interstellar medium (ISM) based on self-consistent two-dimensional
114: fluid simulations. The model describes a partially ionized
115: magnetofluid interstellar medium (ISM) that couples a neutral hydrogen
116: fluid with a plasma primarily through charge exchange interactions.
117: Charge exchange interactions are ubiquitous in warm ISM plasma and the
118: strength of the interaction depends largely on the relative speed
119: between the plasma and the neutral fluid. Unlike small length-scale
120: linear collisional dissipation in a single fluid, charge exchange
121: processes introduce channels that can be effective on a variety of
122: length-scales that depend on the neutral and plasma densities,
123: temperature, relative velocities, charge exchange cross section and
124: the characteristic length scales. We find, from scaling arguments and
125: nonlinear coupled fluid simulations, that charge exchange interactions
126: modify spectral transfer associated with large scale energy containing
127: eddies. Consequently, the warm ISM turbulent cascade rate prolongs
128: spectral transfer amongst inertial range turbulent modes. Turbulent
129: spectra associated with the neutral and plasma ISM fluids are
130: therefore steeper than those predicted by Kolmogorov's phenomenology.
131:
132:
133: \end{abstract}
134:
135: %% Keywords should appear after the \end{abstract} command. The uncommented
136: %% example has been keyed in ApJ style. See the instructions to authors
137: %% for the journal to which you are submitting your paper to determine
138: %% what keyword punctuation is appropriate.
139:
140: \keywords{magnetohydrodynamics: MHD-turbulence-methods:
141: numerical-(Sun:) solar wind-ISM: kinematics and
142: dynamics-interplanetary medium}
143:
144:
145: %% From the front matter, we move on to the body of the paper.
146: %% In the first two sections, notice the use of the natbib \citep
147: %% and \citet commands to identify citations. The citations are
148: %% tied to the reference list via symbolic KEYs. The KEY corresponds
149: %% to the KEY in the \bibitem in the reference list below. We have
150: %% chosen the first three characters of the first author's name plus
151: %% the last two numeral of the year of publication as our KEY for
152: %% each reference.
153:
154:
155: \section{Introduction}
156:
157: The local interstellar medium (LISM) surrounding the heliosphere is
158: warm ($\sim 7000^o K$) and consequently partially ionized. Although
159: the ionization fraction of the LISM is not conclusively established
160: \citep{slavin2006}[see also \citep{muller2006,zank2006}], it is
161: possible that the plasma number density is as small as $0.05 cm^{-3}$
162: and the neutral atomic H number density as much as $0.16
163: cm^{-3}$. Indeed, much of the interstellar medium is warm and
164: partially ionized. In the LISM, the low density plasma and neutral H
165: gas are coupled primarily through the process of charge exchange. On
166: sufficiently large temporal and spatial scales, a partially ionized
167: plasma is typically regarded as equilibrated; this is the case for the
168: LISM. However, the region between the bow shock of our heliosphere and
169: the heliopause \citep{zank1999} is not equilibrated because the charge
170: exchange mean free path (mfp) and the size of the region are
171: comparable. The neutral H distribution in the region is complex
172: \citep{pauls1995,zank1996,zank1996b,jacob2006} possessing an
173: approximately Maxwellian core that is broadened by hot H atoms produced
174: in the inner heliosheath \citep{zank1996b} and fast neutral H produced
175: in the supersonic solar wind \citep{zank1996b,zank1999}. Similarly,
176: the atmospheres of many stars, especially cooler stars, that are
177: embedded in a partially ionized interstellar medium will have outer
178: astrosheaths that are partially ionized. Several examples of such
179: partially ionized astropsheres have been observed using Lyman-$\alpha$
180: absorption measurements \citep{gayley1996, wood2000}.
181:
182:
183: Turbulent fluctuations in a partially ionized LISM are not only
184: potentially important in the context of the global heliospheric ISM
185: interaction, but are also instrumental to our understanding of many
186: astrophysical phenomena including the energization and transport of
187: cosmic rays, gamma-ray bursts, ISM density spectra etc. Besides the
188: ISM, a partially ionized plasma environment represent an important
189: component of the Saturnion magnetosphere, especially that part
190: influenced by mass loss from the moon Encephalus \citep{ip1997}, in
191: Neptune's magnetosphere e.g., \citep{hill1990}, and also in the
192: astrosphere of stars embedded in a cloud of neutral H or even some
193: stellar winds. For magnetosphere, the partially ionized plasma
194: sometimes comprises a mixture of water group and O neutrals and pickup
195: ions and solar wind plasma. The partially ionized plasma environment
196: is also relevant to the study of neutron stars \citep{Potekhin2005}
197: where the plasma and strong magnetic field interaction still remains
198: an unresolved issue.
199:
200:
201: The interaction of a neutral gas and plasma can be dated back to the
202: seminal work by Kulsrud \& Pearce (1969) in the context of cosmic ray
203: propagation, where it was shown that neutral component damps Alfv\'en
204: waves. Neutrals interacting with plasma via a relative drag process
205: results in ambipolar diffusion. Ambipolar diffusion plays a crucial
206: role in the dynamical evolution of the near solar atmosphere,
207: interstellar medium, and molecular clouds and star formation. For
208: instance, Oishi \& Mac Low (2006) investigated the inability of
209: ambipolar diffusion to set a characteristic mass scale in molecular
210: clouds and find that substantial structure persists below the
211: ambipolar diffusion scale because of the propagation of compressive
212: slow mode MHD waves at smaller scales. It was further shown that the
213: spectral behaviour is not influenced by the ambipolar diffusive
214: process unlike viscous damping. Leake et al (2005) showed that the
215: lower chromosphere contains neutral atoms, the existence of which
216: greatly increases the efficiency of wave damping due to collisional
217: friction momentum transfer. They find that Alfv\'en waves with
218: frequencies above 0.6Hz are completely damped and frequencies below
219: 0.01 Hz are unaffected. Khodachenko et al (2006) undertook a
220: quantitative comparative study of the efficiency of the role of
221: (ion-neutral) collisional friction, viscous and thermal conductivity
222: mechanisms in damping MHD waves in different parts of the solar
223: atmosphere. It was pointed out by the authors that a correct
224: description of MHD wave damping requires the consideration of all
225: energy dissipation mechanisms through the inclusion of the
226: appropriate terms in the generalized Ohm’s law, the momentum, energy
227: and induction equations. Molecular clouds are known to be supported,
228: at least in part, by magnetic fields. The removal of magnetic fields
229: thus represents an important component of the star formation
230: process. In the most studied scenario, field removal occurs through
231: the action of ambipolar diffusion, wherein magnetic fields are tied
232: to the ionized component, which drifts relative to the more dominant
233: neutral component of the gas (Mestel \& Spitzer 1956, Mouschovias
234: 1976, Nakano 1979, Shu 1983, Nakano 1984, Lizano \& Shu 1989, Basu \&
235: Mouschovias 1994). The role of magnetic fields and ion-neutral
236: friction in regulating gravitationally driven fragmentation of
237: molecular clouds was studied by Kudoh et al (2007). Mestel \&
238: Spitzer (1956) pointed out that even if clouds are magnetically
239: supported, ambipolar diffusion (resulting from ion-neutral drag) will
240: cause the support to be lost and stars to form. Li \& Nakamura
241: (2004) and Nakamura \& Li (2005) have studied turbulent fragmentation
242: processes for a magnetized sheet including the effect of ion-neutral
243: friction. They find that a mildly subcritical cloud can undergo
244: locally rapid ambipolar diffusion and form multiple fragments because
245: of an initial large scale highly supersonic compression wave. Padoan
246: et al (2000) calculated frictional heating by ion-neutral (or
247: ambipolar) drift in turbulent magnetized molecular clouds and showed
248: that the ambipolar heating rate per unit volume depends on field
249: strength for constant rms Mach number of the flow, and on the
250: Alfv\'enic Mach number. Furthermore, ion-neutral friction, has long
251: been thought to be an important energy dissipation mechanism in
252: molecular clouds, and therefore a significant heating mechanism for
253: molecular cloud gas (Scalo 1977, Goldsmith \& Langer 1978, Zweibel \&
254: Josafatsson 1983, Elmegreen 1985). These are only a few, amongst
255: numerous other, studies that illustrate the importance of a neutral
256: gas component in the dynamics of partially ionized magnetized
257: astrophysical plasmas.
258:
259:
260: On even smaller scales, the edge region of a tokamak (a donut shaped
261: toroidal experimental device designed to achieve thermonuclear fusion
262: reaction in an extremely hot plasma) is partially ionized. In a
263: tokmak, the neutral particles result from effects such as gas puffing,
264: impurity injection, recombination, charge exchange, and possibly
265: neutral beam injection processes. The presence of neutrals can
266: potentially alter the dynamics of zonal flows and cross-field
267: diffusivity. For instance, several experiments have demonstrated that
268: the transition from L-mode (low confinement) to H-mode (high
269: confinement) can be significantly affected by neutral atoms in the
270: edge of a tokamak plasma \citep{singh2004}. The edge region
271: predominantly consists of neutral species resulting from recycling
272: from the wall, and from the limiter and/or divertor plate. These
273: neutrals are present in significant numbers (e.g. $10^{10}- 10^{11}~
274: cm^{-3}$, which is about 5-8\% of the plasma density near the edge)
275: and affects the poloidal momentum balance and the L-H transition
276: threshold. Thus they a play crucial role in regulating transport
277: processes in tokamak fusion plasmas. The neutral gas can interact
278: with the magnetized plasma via charge exchange and other effects.
279: Furthermore, Doublet III-D tokamak (DIII-D) data shows that there is a
280: significant correlation between the power threshold for the L-H
281: transition and the poloidally averaged neutral density at the 95\%
282: flux surface \citep{groebner2001}. To understand the role of neutral
283: gas on H-mode physics, asymmetric gas puffing experiments have also
284: been performed in COMPASS-D and MAST devices
285: \citep{valovie2001,valovie2002}.
286:
287:
288:
289:
290: In all these environments, turbulence is ubiquitous. Turbulence
291: involves the nonlinear coupling and transfer of energy across a
292: multitude of spatial and temporal scales. The coupling of plasma to a
293: neutral gas via charge exchange, especially in a non-equilibrated
294: environment, introduces both a length scale distinct from the usual
295: collisional mfp and alternate channels and mode couplings for the
296: transfer of energy. Certainly, in the local ISM, the charge exchange
297: cross-section of e.g. neutral hydrogen atoms is larger than the
298: collisional cross-section (e.g. Zank 1999), implying that the former
299: governs much of the small-scale physics. Fluctuations with length
300: scales that exceed the charge exchange mean free path $\ell_{ce}$ will
301: obviously be effected by the neutral gas, such as the damping of
302: linear modes (Shaikh \& Zank 2006), but it is less clear how turbulent
303: plasma motion and nonlinear couplings will be mediated. For plasma
304: fluctuations on scales smaller than $\ell_{ce}$, the role of neutral
305: gas is even less clear.
306:
307: Further, it is worth mentioning that the charge exchange process in
308: its simpler (and leading order) form can be treated like a friction or
309: viscous drag term in the fluid momentum equation, describing the
310: relative difference in the ion and neutral fluid velocities. The drag
311: imparted in this manner by a collision between ion and neutral also
312: causes ambipolar diffusion, a mechanism used to describe the Alfv\'en
313: wave damping by cosmic rays (Kulsrud \& Pearce 1969) and also
314: discussed by Oishi \& Mac Low (2006) in the context of molecular
315: clouds. The thermodynamical properties of the latter are significantly
316: different from that found in the heliosheath region. Interestingly,
317: unlike linear damping, ambipolar diffusion does not terminate an
318: isothermal MHD turbulence cascade (Oishi \& Mac Low 2006). Instead, it
319: damps some linear MHD waves. In the context of MHD turbulence, this
320: issue had also been addressed by Cho \& Lazarian (2003) who discussed
321: the viscosity-damped effects on inertial subrange associated with
322: magnetic field fluctuations in compressible (isothermal) MHD
323: turbulence. Ion-neutral friction nonetheless describes a leading order
324: interaction between the ion and neutral fluid species where thermal
325: speeds associated with both the fluids are of little significance or
326: ignorable and plasma phenomena are describable predominantly by
327: isothermal processes. This description is not appropriate to describe
328: the outer heliosheath plasma where proton as well as neutral hydrogen
329: temperatures can be as high as $20,000~^oK$. The thermal speeds
330: associated with the heliosheath protons and neutrals corresponding to
331: such a high temperature cannot therefore be ignored for the following
332: reasons. Firstly, since the thermal speed describes a significant
333: fraction of the energy and is proportional to the relative speeds in
334: the momentum and energy charge exchange sources, it should not be
335: neglected for plasmas that are not cold. It can potentially influence
336: the momentum and energy sources during the course of evolution. This
337: means that the interaction between the heliosheath protons and neutral
338: atoms is not mediated only by the corresponding bulk fluid speeds, but
339: it also depends critically on their thermal speeds. Secondly, for
340: elastic and charge exchange collisions, the initial particle
341: distribution can be well described by Maxwellians even with different
342: streaming velocities and temperatures. Higher order effects are
343: introduced in the dynamics because the distributions are not really
344: Maxwellians due to prior collisions as pointed out by McNutt et
345: al. (1998). The latter introduces Navier-Stokes-like corrections
346: (viscosity and thermal conductivity like terms) to the neutral
347: component of the equation. A consistent treatment of such processes,
348: describing the higher order corrections, in the context of the
349: heliosheath requires that the Maxwellians be multiplied by transfer
350: integrals, as done in Pauls et al (1995). Thirdly, plasma and neutral
351: fluids in the heliosheath region are compressible enough to possess a
352: finite component of thermal energy associated with a non-adiabatic
353: exchange of energy amongst the local fluctuations. Thus in the
354: context of supersonic heliosheath turbulent fluctuations, the
355: characteristic speed of plasma-neutral coupled fluid needs to be
356: treated on equal footing with the thermal speed. Consequently, thermal
357: corrections appear as higher order terms along with the leading order
358: ambipolar-like diffusive forces in the charge exchange expressions as
359: described in section 3. This will be evident from the full energy
360: equation and the subsequent charge exchange energy transfer functions
361: employed in our model.
362:
363:
364: Multiple processes can couple the interstellar medium plasma and
365: neutral gas. Besides charge exchange scattering due to an induced
366: dipole moment in the neutral atom, depending on the temperature
367: (Banks, 1966), is possible. Interestingly enough, variations in
368: cross-sections can also be expected for different ions in the same
369: neutral gas at higher temperatures (Banks, 1966). Further processes,
370: especially in the context of heliospheric processes, are
371: photoionization, electron impact ionization, recombination, and
372: H-H$^+$ Coulomb collisions (see e.g. Zank, 1999). However,
373: cross-section associated with most of them is small compared to that
374: of charge exchange between hydrogen atom and ions in the heliosheath
375: region. We therefore do not incorporate them in our current model.
376:
377:
378: In view of these comments, we have developed a self-consistent
379: plasma-neutral ISM turbulence simulation model based on fluid modeling
380: of neutral and plasma fluids \citep{zank1999,pauls1995,zank1996b}. Our
381: model treats the plasma and neutral species as distinguishable fluids
382: and couples them through charge exchange. The model essentially
383: employs a set of magnetohydrodynamic equations that describe the
384: turbulent plasma motions, whereas the neutrals are described by
385: hydrodynamic equations. A fluid approach for the local interstellar
386: medium is justifiable since the charge exchange and collisional mean
387: free path ensures approximate thermodynamical equipartition between
388: plasma and neutral species with the result that both plasma and
389: neutrals can be described by a Maxwellian distribution. Besides the
390: local ISM, the outer heliosheath can also be described reasonably well
391: by a fluid model of this kind, as was done in the model of
392: \citet{pauls1995}. For a non-equilibrated description, a more
393: elaborate multi-fluid model can be developed, as was done by
394: \citet{zank1996b,zank1996}, and this could be applied to
395: straightforwardly to both the outer heliosheath and very local
396: ISM. This would be more important to the outer heliosheath since a
397: multi-fluid model would then include hot inner heliosheath neutral
398: atoms deposited through secondary charge exchange in the outer
399: heliosheath \citep{zank1996b,zank1996,zank1999}. The schematic in
400: \Fig{fig0} depicts the simulation regions in which we are particularly
401: interested i.e. the outer heliosheath and the local interstellar
402: medium. The importance of ISM turbulence is clearly evident from this
403: figure in the context of cosmic ray modulation, large-scale
404: heliospheric structure and particle acceleration.
405:
406: In this paper, we focus on the multi-scale evolution of a partially
407: ionized plasma, comprising multiple fluids, where each fluid evolves
408: under the influence of the other through a complex interaction
409: process. We investigate the energy cascade between LISM
410: turbulent-fluctuations in a partially ionized plasma where the plasma
411: and neutral gas are coupled through charge exchange. Here, we, present
412: a numerical simulation of ISM turbulence that self-consistently
413: evolves both the plasma and neutral fluids when coupled by charge
414: exchange. We restrict our attention to a neutral hydrogen (H) gas, so
415: that plasma and neutral densities are conserved by charge exchange
416: interactions. Hence the corresponding continuity equations do not
417: include charge exchange terms. One of the points that emerges from
418: our coupled multi fluid LISM simulations is that neutrals enhance
419: turbulent cascade rates and lead to much steeper spectra compared to
420: that predicted by the usual Kolmogorov theory \citep{k41} for pure MHD
421: or hydrodynamics \citep{shaikh2007}. In Section 2, we discuss the
422: equations of a coupled plasma-neutral ISM turbulence model, their
423: validity, the underlying assumptions and the normalizations. Section 3
424: deals with a quantitative derivation of charge exchange source terms,
425: associated with the plasma and neutral fluids, based on the
426: time-dependent Boltzmann equation. Section 4 describes the results of
427: our nonlinear, coupled, self-consistent ISM fluid simulations. It is
428: shown, in section 5, that the coupling of plasma and neutral fluids
429: leads to an enhanced spectral transfer in the inertial range, which
430: results in a steeper ISM turbulent spectrum than predicted by the
431: Kolmogorov theory. Finally, conclusions are presented in Section 6.
432:
433:
434: \section{Model Equations for a partially ionized LISM}
435: The assumptions that are intrinsic to our model of turbulence in a
436: partially ionized ISM are the following. (i) Fluctuations in the
437: plasma and neutral fluids are isotropic, homogeneous, thermally
438: equilibrated and turbulent, and (ii) Neither a mean magnetic field nor
439: velocity flows are present initially. Local mean flows may
440: subsequently be generated by self-consistently excited nonlinear
441: instabilities. (iii) The characteristic turbulent correlation
442: length-scales ($\lambda_c \sim 1/k_c$) are typically bigger than
443: charge-exchange mean free path lengths ($\lambda_{ce}\sim 1/k_{ce}$)
444: in the ISM flows, i.e $\lambda_c\gg \lambda_{ce}$ or $k_{ce}/k_c \gg
445: 1$. The latter inequality is also consistent with
446: \citet{Florinski2003,Florinski2005}. Nevertheless, they are large
447: enough to treat any localized shocks as smooth discontinuities. In
448: other words, the characteristic shock length-scales are small compared
449: to the ISM turbulent fluctuation length-scales, and finally (iv)
450: boundary conditions are periodic, essentially a box of interstellar
451: plasma, which is a natural and most appropriate choice for modeling
452: turbulence in the local ISM.
453:
454: While most of the above assumptions are appropriate to realistic ISM
455: turbulent flows, so allowing us to use MHD and hydrodynamic
456: descriptions for the plasma and the neutral components respectively,
457: the latter has been criticized in the context of heliospheric flows
458: within the heliopause owing to the large charge-exchange mean free
459: path (about the size or even bigger than the extent of the
460: heliosphere). The use of a fluid description for neutrals in the
461: inner heliosheath is therefore inappropriate. In the LISM and outer
462: heliosheath, we are nonetheless not restricted by a large mfp because
463: the plasma and neutral fluid remain close to thermal equilibirium and
464: behave as Maxwellian fluids. Our model simulates the local ISM and
465: outer heliosheath. The fluid model describing nonlinear turbulent
466: processes in the interstellar medium, in the presence of charge
467: exchange, can be cast into plasma density ($\rho_p$), velocity (${\bf
468: U}_p$), magnetic field (${\bf B}$), pressure ($P_p$) components
469: according to the conservative form
470: \be
471: \label{mhd}
472: \frac{\partial {\bf F}_p}{\partial t} + \nabla \cdot {\bf Q}_p={\cal Q}_{p,n},
473: \ee
474: where,
475: \[{\bf F}_p=
476: \left[
477: \begin{array}{c}
478: \rho_p \\
479: \rho_p {\bf U}_p \\
480: {\bf B} \\
481: e_p
482: \end{array}
483: \right],
484: {\bf Q}_p=
485: \left[
486: \begin{array}{c}
487: \rho_p {\bf U}_p \\
488: \rho_p {\bf U}_p {\bf U}_p+ \frac{P_p}{\gamma-1}+\frac{B^2}{8\pi}-{\bf B}{\bf B} \\
489: {\bf U}_p{\bf B} -{\bf B}{\bf U}_p\\
490: e_p{\bf U}_p
491: -{\bf B}({\bf U}_p \cdot {\bf B})
492: \end{array}
493: \right],\\
494: {\cal Q}_{p,n}=
495: \left[
496: \begin{array}{c}
497: 0 \\
498: {\bf Q}_M({\bf U}_p,{\bf V}_n, \rho_p, \rho_n, T_n, T_p) \\
499: 0 \\
500: Q_E({\bf U}_p,{\bf V}_n,\rho_p, \rho_n, T_n, T_p)
501: \end{array}
502: \right]
503: \]
504: and
505: \[ e_p=\frac{1}{2}\rho_p U_p^2 + \frac{P_p}{\gamma-1}+\frac{B^2}{8\pi}.\]
506: The above set of plasma equations is supplemented by $\nabla \cdot {\bf
507: B}=0$ and is coupled self-consistently to the ISM neutral density
508: ($\rho_n$), velocity (${\bf V}_n$) and pressure ($P_n$) through a set
509: of hydrodynamic fluid equations,
510: \be
511: \label{hd}
512: \frac{\partial {\bf F}_n}{\partial t} + \nabla \cdot {\bf Q}_n={\cal Q}_{n,p},
513: \ee
514: where,
515: \[{\bf F}_n=
516: \left[
517: \begin{array}{c}
518: \rho_n \\
519: \rho_n {\bf V}_n \\
520: e_n
521: \end{array}
522: \right],
523: {\bf Q}_n=
524: \left[
525: \begin{array}{c}
526: \rho_n {\bf V}_n \\
527: \rho_n {\bf V}_n {\bf V}_n+ \frac{P_n}{\gamma-1} \\
528: e_n{\bf V}_n
529: \end{array}
530: \right],\\
531: {\cal Q}_{n,p}=
532: \left[
533: \begin{array}{c}
534: 0 \\
535: {\bf Q}_M({\bf V}_n,{\bf U}_p, \rho_p, \rho_n, T_n, T_p) \\
536: Q_E({\bf V}_n,{\bf U}_p,\rho_p, \rho_n, T_n, T_p)
537: \end{array}
538: \right],
539: \]
540: \[e_n= \frac{1}{2}\rho_n V_n^2 + \frac{P_n}{\gamma-1}.\]
541:
542: Equations (\ref{mhd}) to (\ref{hd}) form an entirely self-consistent
543: description of the coupled ISM plasma-neutral turbulent fluid.
544: Several points are worth noting. The charge-exchange momentum sources
545: in the plasma and the neutral fluids, i.e. Eqs. (\ref{mhd}) and
546: (\ref{hd}), are described respectively by terms ${\bf Q}_M({\bf
547: U}_p,{\bf V}_n,\rho_p, \rho_n, T_n, T_p)$ and ${\bf Q}_M({\bf
548: V}_n,{\bf U}_p,\rho_p, \rho_n, T_n, T_p)$. A swapping of the plasma
549: and the neutral fluid velocities in this representation corresponds,
550: for instance, to momentum changes (i.e. gain or loss) in the plasma
551: fluid as a result of charge exchange with the ISM neutral atoms
552: (i.e. ${\bf Q}_M({\bf U}_p,{\bf V}_n,\rho_p, \rho_n, T_n, T_p)$ in
553: Eq. (\ref{mhd})). Similarly, momentum change in the neutral fluid by
554: virtue of charge exchange with the plasma ions is indicated by ${\bf
555: Q}_M({\bf V}_n,{\bf U}_p,\rho_p, \rho_n, T_n, T_p)$ in
556: Eq. (\ref{hd}). In the absence of charge exchange interactions, the
557: plasma and the neutral fluid are de-coupled trivially and behave as
558: ideal fluids. While the charge-exchange interactions modify the
559: momentum and the energy of plasma and the neutral fluids, they
560: conserve density in both the fluids (since we neglect photoionization
561: and recombination). Nonetheless, the volume integrated energy and the
562: density of the entire coupled system will remain conserved in a
563: statistical manner. The conservation processes can however be altered
564: dramatically in the presence of any external forces. These can include
565: large-scale random driving of turbulence due to any external forces or
566: instabilities, supernova explosions, stellar winds, etc. Finally, the
567: magnetic field evolution is governed by the usual induction equation,
568: i.e. Eq. (\ref{mhd}), that obeys the frozen-in-field theorem unless
569: some nonlinear dissipative mechanism introduces small-scale damping.
570:
571:
572: The underlying ISM turbulence model can be non-dimensionalized
573: straightforwardly using a typical ISM scale-length ($\ell_0$), density
574: ($\rho_0$) and velocity ($v_0$). The normalized plasma density,
575: velocity, energy and the magnetic field are respectively;
576: $\bar{\rho}_p = \rho_p/\rho_0, \bar{\bf U}_p={\bf U}_p/v_0,
577: \bar{P}_p=P_p/\rho_0v_0^2, \bar{\bf B}={\bf B}/v_0\sqrt{\rho_0}$. The
578: corresponding neutral fluid quantities are $\bar{\rho}_n =
579: \rho_n/\rho_0, \bar{\bf U}_n={\bf U}_n/v_0,
580: \bar{P}_n=P_n/\rho_0v_0^2$. The momentum and the energy
581: charge-exchange terms, in the normalized form, are respectively
582: $\bar{\bf Q}_m={\bf Q}_m \ell_0/\rho_0v_0^2, \bar{Q}_e=Q_e
583: \ell_0/\rho_0v_0^3$. The non-dimensional temporal and spatial
584: length-scales are $\bar{t}=tv_0/\ell_0, \bar{\bf x}={\bf
585: x}/\ell_0$. Note that we have removed bars from the set of
586: normalized coupled ISM model equations (\ref{mhd}) \& (\ref{hd}). The
587: charge-exchange cross-section parameter ($\sigma$), which does not
588: appear directly in the above set of equations (see the subsequent
589: section for more detail), is normalized as $\bar{\sigma}=n_0 \ell_0
590: \sigma$, where the factor $n_0\ell_0$ has dimension of (area)$^{-1}$.
591: By defining $n_0, \ell_0$ through
592: $\sigma_{ce}=1/n_0\ell_0=k_{ce}^{-2}$, we see that there exists a
593: critical charge exchange wavenumber ($k_{ce}$) associated with the
594: coupled ISM plasma-neutral turbulent system. For a characteristic
595: density, this corresponds physically to an area defined by the charge
596: exchange mode being equal to (mfp)$^2$. Thus the larger the area, the
597: higher is the probability of charge exchange between plasma ions and
598: neutral atoms, as illustrated in Fig. (2). To be precise, `$k_{ce}$'
599: is a (lengthscale)$^{-1}$ that typically helps us determine whether or
600: not a particular turbulent fluctuation length scale undergoes charge
601: exchange in our model. It is strictly in that sense, we refer to it as
602: a critical wavenumber. Clearly, there exist three different regimes
603: depending on whether (i) $k<k_{ce}$, (ii) $k \simeq k_{ce}$ or (iii)
604: $k>k_{ce}$. In the heliosheath, the probability that charge exchange
605: can directly modify those modes satisfying $k<k_{ce}$ is high compared
606: to modes satisfying $k>k_{ce}$. Since the charge exchange
607: length-scales are much smaller than the ISM turbulent correlation
608: scales, this further allows many charge exchange interactions amongst
609: the nonlinear turbulent ISM modes before they cascade energy in one
610: unit convective time. This is illustrated schematically in Fig. (2).
611:
612: It is interesting to compare our model with other work. Of particular
613: relevance is the classic work by Kulsrud and Pearce (1969) who
614: investigated the interaction of galactic comic rays and Alfi\'en waves
615: in the interstellar medium and also considered ion-neutral collision
616: as a damping mechanism for Alfv\'en waves. Ion-neutral collisions are
617: also reported by McIvor (1977) to be a dominant mechanism for damping
618: interstellar medium turbulence. The damping (due to ion-neutral
619: collisions) rates for Alfv\'en waves estimated by Kulsrud and Pearce
620: (1969) are based on $\nu_0 \simeq n_0 v_{th} \sigma$, where $\nu_0,
621: n_0, v_{th}$ and $\sigma$ are respectively the ion-neutral collision
622: frequency, neutral density, thermal speed and collision cross-section.
623: The damping wavenumber associated with this frequency can be estimated
624: to be $k\simeq n_0 v_{th} \sigma/V_A$. The latter determines
625: essentially the dissipation wavenumber associated with the damping of
626: small scale Alfv\'en waves in ISM turbulence (Kulsrud and Pearce
627: 1969). The turbulent cascade is further expected to be terminated
628: beyond this wavenumber. It is clear from this expression that the
629: higher the thermal speed, the larger the dissipation
630: wavenumber. Correspondingly in real space, dissipation will be
631: concentrated on smaller length-scales for higher thermal
632: speeds. Furthermore, the thermal speed is also proportional to the
633: temperature. Thus, given the heliosheath parameters, $v_{Thu}$
634: associated with {\it heliosheath} neutral atoms will be at least two
635: orders of magnitude larger compared with that results from Eq. (47) in
636: Kulsrud and Pearce (1969). This will correspond to a larger
637: dissipative wavenumber in the heliosheath region and essentially means
638: that charge exchange interactions are restricted to only those eddies
639: whose length-scales are comparable to the dissipative length-scales of
640: turbulence. It is not clear if this scenario holds in inner/outer
641: heliosheath turbulence where charge exchange interactions, in
642: principle, can modify any length scale in the inertial range. Part of
643: the reason resides with the fact that charge exchange modifications
644: are proportional to the relative speed between ion and neutral fluids
645: and corresponding densities which can obviously alter the nonlinear
646: cascades by influencing the combined momenta and energies of both the
647: fluids in a self-consistent and subtle manner. Additionally, Kulsrud
648: and Pearce (1969), as does McIvor (1977), assumed a fixed neutral
649: density while computing the damping rates associated with Alfv\'en
650: waves in ISM turbulence. While a fixed neutral density can modify the
651: linear growth rates of the underlying waves, it does not itself evolve
652: and hence no back reaction on Alfv\'en waves and vice versa can be
653: expected. By contrast, our model self-consistently evolves neutral
654: and plasma and describes the mutual feedback of one species on the
655: other. This not only alters the linear interaction process, but it
656: also influences the nonlinear coupling in a subtle manner not readily
657: describable by means of any linear analytic theory. A self-consistent
658: coupling is absolutely essential to understand nonlinear turbulent
659: cascades in the inner/outer heliosheath. It is to be noted further
660: that we do not attempt to investigate damping of heliosheath
661: turbulence by Alfv\'en waves that experience ion-neutral collisions,
662: unlike McIvor (1977). While Alfv\'en waves are intrinsically present
663: in our model, turbulent cascades are least affected by them along the
664: mean magnetic field (Shebalin \& Montgomery 1983). Alfv\'en waves can
665: be crucial in providing a local anisotropy in the spectral transfer
666: due to a mean or local field though. However, our prime focus here is
667: to develop an understanding of heliosheath turbulence where plasma and
668: neutral fluids evolve through a mutual interaction mediated by charge
669: exchange. As evident from the complexity associated with these
670: interactions, nonlinear turbulent cascades are likely to be affected
671: at any length scale in the heliosheath inertial range, unlike a linear
672: dissipative processes.
673:
674:
675: In the subsequent section, we derive an exact quantitative form of the
676: sources due to charge exchange in ISM turbulence.
677:
678:
679:
680: \section{Charge Exchange Sources}
681: The charge exchange terms can be obtained from the Boltzmann transport
682: equation that describes the evolution of a neutral distribution
683: function $f_n=f({\bf x}, v_x, v_y, v_z, t)$ in a six-dimensional phase
684: space defined respectively by position and velocity vectors $({\bf x},
685: v_x, v_y, v_z)$ at each time $t$. Here we follow \citet{pauls1995} in
686: computing the charge exchange terms from various moments of the
687: Boltzmann equation. The Boltzmann equation for the neutral
688: distribution contains a source term proportional to the proton
689: distribution function $f_p$ and a loss term proportional to the
690: neutral distribution function $f_n$, \eqa
691: \label{bol}
692: \frac{\partial f_n}{\partial t} + {\bf v}_n \cdot \nabla f_n +
693: \frac{\bf F}{m} \cdot \nabla_{{\bf v}_n} f_n =
694: f_{p}({\bf x}, v_x, v_y, v_z, t) \int f_n({\bf x}, v_x, v_y, v_z, t) |{ v}_n- {u}_p|
695: \sig(v_{rel}) d^3{v}_n \nonumber \\
696: -f_n({\bf x}, v_x, v_y, v_z, t) \int
697: f_p({\bf x}, v_x, v_y, v_z, t) |{ u}_p- { v}_n| \sig(v_{rel})
698: d^3{ u}_p. \eeq Here $f_p, ~{u}_p$ represent respectively the
699: proton distribution function and velocity. $\sig$ is the charge
700: exchange cross-section (between neutrals and plasma protons), $m$ is
701: the mass of particle, and ${\bf F}$ represents forces acting on the
702: fluid. The charge exchange parameter has a logarithmically weak
703: dependence on the relative speed ($v_{rel}=|{u}_p- {v}_n|$) of
704: the neutrals and the protons through $\sig = [(2.1-0.092 \ln
705: (v_{rel})) 10^{-7} cm]^2$ \citep{fite}. This cross-section is valid
706: as long as energy does not exceed $1eV$, which usually is the case in
707: the inner/outer heliosphere. Beyond $1eV$ energy, this cross-section
708: yields a higher neutral density. This issue is not applicable to our
709: model and hence we will not consider it here. The density, momentum,
710: and energy of the thermally equilibrated Maxwellian proton and neutral
711: fluids can be computed from \eq{bol} by using the zeroth, first and
712: second moments $\int f_{\xi} d^3\xi, \int m{\bf \xi} f_{\xi} d^3\xi$
713: and $\int m\xi^2/2 f_{\xi} d^3\xi$ respectively, where $\xi={u}_p$
714: or ${v}_n$. Since charge exchange conserves the density of the
715: proton and neutral fluids, there are no sources in the corresponding
716: continuity equations. We, therefore, need not compute the zeroth
717: moment of the distribution function. Computing directly the first
718: moment from \eq{bol}, we obtain the neutral fluid momentum equation as
719: given by \eq{hd}. The entire rhs of \eq{bol} can now be replaced by a
720: momentum transfer function ${\bf Q}_M({v}_n,{u}_p) $ which
721: reads \be
722: \label{qm}
723: {\bf Q}_M({v}_n,{u}_p) = \bar{\mu}({u}_p,{v}_n)-\bar{\mu}({v}_n,{u}_p),
724: \ee
725: where ${\bf Q}_M$ and $ \bar{\mu}$, the transfer integral, are
726: vector quantities. The transfer integrals describe the
727: charge exchange transfer of momentum from proton to neutral fluid and
728: vice versa. The first term on the rhs of
729: Eq. (4) can be expressed by
730: \[\bar{\mu}({u}_p,{v}_n) = f_p({\bf x}, v_x,v_y,v_z,t)\beta({u}_p,{v}_n)\]
731: where,
732: \[\beta({u}_p,{v}_n)=\int f_n({\bf x}, v_x,v_y,v_z,t) |{v}_n- {u}_p|
733: \sig(v_{rel}) d^3{v}_n.\]
734: Considering a Maxwellian distribution for the neutral atoms, we simplify
735: $\beta({u}_p,{v}_n)$ as
736: follows,
737: \[\beta({u}_p,{v}_n)= \sig n_n V_{T_n}
738: \sqrt{\frac{4}{\pi}+\frac{({v}_n- {u}_p)^2}{V_{T_n}^2}}.\] Note that
739: the above expression emerges directly from a straightforward
740: integration of sources in the rhs of the Boltzmann Eq. (3). To obtain
741: the expression for the momentum transferred from proton to neutral (or
742: vice versa), we need to take a second moment of the
743: $\bar{\mu}({u}_p,{v}_n)$ expression. This is shown in the following,
744: \[\bar{\mu}({u}_p,{v}_n) = m {\bf v}_n I_0({u}_p,{v}_n) + m ({\bf u}_p-{\bf v}_n) I_1({u}_p,{v}_n).\]
745: where $I_0$ and $I_1$ are transfer integrals that can be written as follows,
746: \[I_0({u}_p,{v}_n)= \int f_p({\bf x}, v_x,v_y,v_z,t)\beta({u}_p,{v}_n) ~d^3{u}_p;\]
747: \[I_1({u}_p,{v}_n)= \int {v}_n f_p({\bf x}, v_x,v_y,v_z,t)\beta({u}_p,{v}_n) ~d^3{u}_p.\]
748: Assuming a Maxwellian distribution for plasma protons and using
749: $\beta({u}_p,{v}_n)$ from the above expression, we can
750: straightforwardly evaluate the transfer integrals $I_0$ and $I_1$ (see
751: Pauls et al (1995) for details). We further write the
752: complete form of the first term on the rhs of Eq. (4) as follows,
753: \[\bar{\mu}({u}_p,{v}_n) = m\sig n_p n_n \left[U_{{u}_p,{v}_n}^\ast{\bf v}_n - ({\bf u}_p-{\bf v}_n)
754: \frac{ V_{T_n}^2}{\delta V_{{u}_p,{v}_n}} \right].\] In a similar
755: manner, we can evaluate the second term on the rhs of Eq. (4), which
756: yields the following form,
757: \[\bar{\mu}({v}_n,{u}_p) = m\sig n_{p}n_n \left[U_{{v}_n,{u}_p}^\ast {\bf u}_p - ({\bf v}_n-{\bf u}_p) \frac{V_{T_p}^2}{\delta V_{{v}_n,{u}_p}} \right],\]
758: where $\delta V_{{u}_p,{v}_n} =
759: \left[4\left(\frac{4}{\pi}V_{T_p}^2+\Delta U^2 \right)
760: +\frac{9\pi}{4}V_{T_n}^2 \right]^{1/2}, \delta V_{{v}_n,{u}_p} = \left[4\left(\frac{4}{\pi}V_{T_n}^2+\Delta U^2 \right)
761: +\frac{9\pi}{4}V_{T_{n}}^2 \right]^{1/2}$ and $U^\ast=U_{{u}_p,{v}_n}^\ast = U_{{v}_n,{u}_p}^\ast =
762: \left[\frac{4}{\pi}V_{T_{p}}^2+\frac{4}{\pi}V_{T_{n}}^2 +\Delta
763: U^2\right]^{1/2}, \Delta {u} = {u}_p-{v}_n$ \citep{pauls1995}. On
764: substituting these expressions in the momentum transfer function, we
765: obtain
766: \be
767: \label{qm2}
768: {\bf Q}_M({v}_n,{u}_p) = m\sig n_{p}n_n ({\bf v}_n -{\bf
769: u}_p) \left[ U^\ast + \frac{V_{T_n}^2}{\delta V_{{u}_p,{v}_n}} -\frac{V_{T_p}^2}{\delta V_{{v}_n,{u}_p}}
770: \right].
771: \ee
772: \Eq{qm2} together with \eq{hd} yields the momentum equation for
773: the neutral gas. Swapping the plasma and neutral fluid velocities
774: yields the corresponding source term for the proton fluid momentum
775: equation. The gain or the loss in neutral or proton fluid momentum
776: depends upon the charge exchange sources, which depend upon the
777: relative speeds between neutrals and the protons. The thermal speeds
778: of proton and neutral gas in \eq{qm2} are given respectively by $
779: V_{T_p}^2 = 2K_BT_{p}/m$ (the factor 2 arises because of thermal
780: equilibration in that it is assumed that the temperature of the plasma
781: electrons and protons are nearly identical so that $T_p = T_e + T_{\rm
782: proton} \simeq 2T_p$) and $V_{T_n}^2 = K_BT_{n}/m$. The
783: corresponding temperatures are related to the pressures by $
784: P_{p}=2n_{p}K_BT_{p}$ and $P_{n}=n_{n}K_BT_{n}$, where $n_n, T_n,
785: n_{p}, T_{p}$ are respectively the neutral and plasma density and the
786: temperature, and $K_B$ is the Boltzmann constant.
787:
788:
789: The moment, $\int m\xi^2/2 f_{\xi} d^3\xi$, of the Boltzmann
790: \eq{bol} yields an energy equation for the neutral fluid whose
791: rhs contains the charge exchange energy transfer function
792: \[Q_E({v}_n,{u}_p) = \eta({u}_p,{v}_n)-\eta({v}_n,{u}_p),\]
793: where $\eta({u}_p,{v}_n), ~\eta({v}_n,{u}_p)$ are the
794: energy transfer (from neutral to proton and vice versa) rates. These
795: functions can be computed as follows:
796: \[\eta({u}_p,{v}_n) = \frac{1}{2}
797: mV_n^2 \sig n_{p}n_n U^\ast + \frac{3}{4}m V_{T_n}^2 \sig n_{p}n_n \Delta
798: V_{{u}_p,{v}_n}- m \sig n_pn_n {\bf v}_n\cdot ({\bf u}_p-{\bf
799: v}_n) \frac{V_{T_n}^2}{\delta V_{{u}_p,{v}_n}}\]
800: and
801: \[\eta({v}_n,{u}_p) = \frac{1}{2}mV_{n}^2 \sigma n_{p}n_n U^\ast + \frac{3}{4}m
802: V_{T_{p}}^2 \sig n_{p}n_n \Delta V_{{v}_n,{u}_p}- m \sig
803: n_{p}n_n {\bf u}_p \cdot ({\bf v}_n-{\bf u}_p) \frac{V_{T_{p}^2}}{\delta
804: V_{{v}_n,{u}_p}}.\]
805: The total energy transfer from neutral to proton fluid, due to charge
806: exchange, can then be written as,
807: \eqa
808: \label{qe}
809: Q_E({v}_n,{u}_p) = \frac{1}{2}m \sig n_{p}n_n U^\ast (V_n^2-U_p^2)+
810: \frac{3}{4}m \sig n_{p}n_n (V_{T_{n}}^2\Delta V_{{u}_p,{v}_n}-V_{T_{p}}^2
811: \Delta V_{{v}_n,{u}_p})
812: \nonumber \\
813: -m \sig n_{p}n_n
814: \left[ {\bf v}_n \cdot ({\bf u}_p-{\bf v}_n)\frac{V_{T_n}^2}{\delta V_{{u}_p,{v}_n}}-
815: {\bf u}_p \cdot ({\bf v}_n-{\bf u}_p)\frac{V_{T_{p}^2}}{\delta V_{{v}_n,{u}_p}}\right],
816: \eeq
817: with $\Delta V_{{u}_p,{v}_n} =
818: [\frac{4}{\pi}V_{T_p}^2+\frac{64}{9\pi}V_{T_n}^2 +\Delta U^2]^{1/2} ~{\rm and} ~
819: \Delta V_{{v}_n,{u}_p} = [\frac{4}{\pi}V_{T_n}^2+
820: \frac{64}{9\pi}V_{T_{p}}^2 +\Delta U^2]^{1/2}$.
821: A similar expression for the energy transfer charge exchange source
822: term of plasma energy in \eq{mhd} can be obtained by exchanging the
823: plasma and neutral fluid velocities. In the normalized momentum and
824: energy charge exchange source terms, the factor $m\sig$ in
825: \eqs{qm2}{qe} is simply replaced by $\bar{\sigma}$.
826:
827:
828:
829: \section{ Nonlinear Simulation of Energy Cascades}
830: It is clear from previous sections that the interaction of the
831: neutrals with a plasma introduces characteristic length and time scales
832: that can separate characteristic time and length scales in ISM
833: turbulence. For instance, the plasma-neutral fluid coupling in the ISM
834: fluctuations introduces charge exchange modes, $k_{ce}$, that are
835: distinctively different from the characteristic turbulent mode
836: $k$. Typically, characteristic turbulent length-scales ($\lambda_c\sim
837: 1/k$) are much bigger than the charge exchange length-scales
838: $\lambda_{ce}\sim1/k_{ce}$, since for example, $\lambda_c \sim
839: (100-1000)\lambda_{ce}$ in the ISM cloud
840: \citep{Florinski2003,Florinski2005}. This essentially corresponds to
841: $k_{ce}/k\gg1$ \citep{zank1999,Florinski2003,Florinski2005}. It should
842: be noted that the charge exchange sources in the neutral and the
843: plasma fluids do not merely damp the smallest scale fluctuations, such
844: as is the case with linear collisional dissipation in turbulent fluid
845: flows, but they can alter the nonlinear cascade rate dramatically
846: \citep{shaikh2006,shaikh2007}. Because of its complex nonlinear
847: character (see \eqs{qm2}{qe}), charge exchange can be {\it effective on
848: a variety of turbulent length-scales} unlike linear dissipation which
849: is operative on small length-scales only. The effect and influence of
850: the charge exchange source terms on the turbulent inertial range of
851: coupled neutral and plasma fluids is therefore a subtle issue and is
852: not restricted only to the damping of small scales in partially
853: ionized, magnetized ISM turbulence. With this in mind, we carry out
854: two different sets of simulations of \eqs{mhd}{hd}, to discern the
855: effects of charge exchange on the ISM plasma-neutral system. In the
856: first set of simulations, \eqs{mhd}{hd} are decoupled (i.e. no charge
857: exchange coupling is assumed) and each evolves as an ideal fluid,
858: whereas in the second set the coupling is included self-consistently.
859: The former provides a reference simulation against which to
860: investigate the coupled system.
861:
862: A two-dimensional (2D) nonlinear fluid code was developed to
863: numerically integrate \eqsto{mhd}{hd}. The 2D simulations are not only
864: computationally less expensive (compared to a fully 3D calculation),
865: but they offer significantly higher resolution (to compute inertial
866: range turbulence spectra) even on moderately-sized clusters like our
867: local Beowulf. The spatial discretization in our code uses a discrete
868: Fourier representation of turbulent fluctuations based on a
869: pseudospectral method, while we use a Runge Kutta 4 method for the
870: temporal integration. All the fluctuations are initialized
871: isotropically (no mean fields are assumed) with random phases and
872: amplitudes in Fourier space. This algorithm ensures conservation of
873: total energy and mean fluid density per unit time in the absence of
874: charge exchange and external random forcing. Additionally, $\nabla
875: \cdot {\bf B}=0$ is satisfied at each time step. Our code is
876: massively parallelized using Message Passing Interface (MPI) libraries
877: to facilitate higher resolution. The initial isotropic turbulent
878: spectrum of fluctuations is chosen to be close to $k^{-2}$ with random
879: phases in both $x$ and $y$ directions. The choice of such (or even a
880: flatter than -2) spectrum does not influence the dynamical evolution
881: as the final state in our simulations progresses towards fully
882: developed turbulence. While the ISM turbulence code is evolved with
883: time steps resolved self-consistently by the coupled fluid motions,
884: the nonlinear interaction time scales associated with the plasma
885: $1/{\bf k} \cdot {\bf U}_p({\bf k})$ and the neutral $1/{\bf k} \cdot
886: {\bf V}_n({\bf k})$ fluids can obviously be disparate. Accordingly,
887: turbulent transport of energy in the plasma and the neutral ISM fluids
888: takes place on distinctively separate time scales.
889:
890: Because of the different nonlinear time scales associated with the
891: plasma and neutral fluids, mode structures can be different in the two
892: fluids when they are evolved together and in isolation
893: (i.e. decoupled). The former is shown in \fig{fig3}{fig4}. The figures
894: respectively show the evolution of various physical quantities in the
895: plasma-neutral coupled system for a typical set of ISM parameters as
896: described in the captions (for typical ISM parameters, see also
897: \cite{zank1999,zank1996b,jacob2006,Florinski2003,Florinski2005}). In
898: the absence of charge exchange (see \Fig{fig3}, right column), the
899: plasma fluid evolves as an ideal MHD fluid and is known to typically
900: develop sheet-like structures in the magnetic field
901: \citep{biskamp}. By contrast, the evolution of the magnetic field is
902: modified substantially when the two fluids are coupled through charge
903: exchange. For instance, the sheet-like structures present in the
904: turbulent magnetic field of the uncoupled plasma system instead are
905: smeared in the coupled plasma-neutral system on the typical
906: sheet-formation time-scales of ideal MHD. The neutral fluid, under
907: the action of charge exchange terms, tends to modify the cascade rates
908: by isotropizing the ISM turbulence on a slower time scale. This point
909: is further elucidated in the subsequent section based on
910: Kolmogorov-like phenomenological arguments. However if the coupled
911: system is evolved further, sheets do form eventually in the magnetic
912: field (after a relatively long time compared to the ideal MHD
913: time-scales) and a turbulent equipartition is set up in the plasma and
914: the neutral fluid modes. In any event, the small-scale sheet-like
915: structures in magnetic field compress (or pinch) the plasma
916: density. Hence the density fluctuations develop identical (thinner
917: than) sheet-like structures that co-exist with small-scale turbulent
918: fluctuations in their spectrum as shown in \Fig{fig3}. The neutral
919: fluid, on the other hand, evolves isotropically as stated above by
920: forming relatively large-scale structures (see \Fig{fig4}).
921:
922:
923:
924:
925:
926: \section{Energy Spectra}
927: Spectral transfer in ISM turbulence progresses under the action of
928: nonlinear interactions as well as charge exchange sources. Energy
929: cascades amongst turbulent eddies of various scale sizes and between
930: the plasma and the neutral fluids. In a freely decaying case, plasma
931: and neutral fluids evolve under the influence of charge exchange
932: forces which dramatically affect the energy cascades in the inertial
933: range. This is evident from \fig{fig5}{fig6} where Kolmogorov-like
934: fully developed turbulent spectra in the inertial range are shown
935: respectively for the coupled neutral and plasma fluids. The neutral
936: fluid energy spectrum in the inertial regime for the coupled system
937: exhibits a $k^{-\alpha}$ spectrum, where the spectral index $\alpha
938: \approx 3.2$. On the other hand, the spectral index for plasma
939: magnetic and kinetic energy spectra for the coupled system is $\alpha
940: \approx 2.4$. The inertial range spectral indices for the coupled
941: plasma-neutral system in our simulations show a {\it significant}
942: deviation from their corresponding uncoupled analogues which are
943: respectively $-3$ and $-1.7$ for the neutral and plasma fluids, also
944: plotted in \fig{fig5}{fig6}. The inertial range turbulent spectra are
945: much steeper for the coupled ISM plasma-neutral system than the
946: corresponding uncoupled or isolated neutral and plasma fluids in the
947: decaying turbulence regime. Consequently, the steep turbulent spectra
948: in the coupled ISM plasma-neutral system gives rise to an enhanced
949: spectral transfer amongst the inertial range modes, and the energy
950: cascade rates associated with the plasma and neutral (coupled) fluids
951: can be understood by invoking a Kolmogorov-like phenomenology as
952: described in the following.
953:
954: The typical nonlinear interaction time-scale in ordinary
955: (i.e. uncoupled) plasma and/or neutral turbulence is given by
956: \be
957: \tau_{nl} \sim \frac{\ell_0}{v_\ell} \sim (kv_k)^{-1},
958: \ee
959: where $v_k$ or $v_\ell$ is the velocity of turbulent eddies. In the
960: presence of charge exchange interactions, the ordinary nonlinear
961: interaction time-scale of fluid turbulence is modified by a factor
962: $k_{ce}/k$ such that the new nonlinear interaction time-scale of ISM
963: turbulence is now
964: \be
965: \tau_{NL} \sim \frac{k_{ce}}{k} \frac{1}{kv_k}.
966: \ee
967: The main justification of using the above expression originates from
968: the fact that turbulent cascade is determined typically by the
969: nonlinear interaction time ($\tau_{nl} \sim 1/kv_k$, where $\tau_{nl}$
970: is the eddy interaction time, $k$ is characteristic turbulent
971: wavenumber, and $v_k$ is characteristic speed of eddy in wavenumber
972: space) associated with the convection of turbulent eddies in the fluid
973: momentum equation. Note that this time scale follows from Kolmogorov's
974: phenomenology (Kolmogorov 1941) and corresponds essentially to the
975: eddy interaction time in an ideal, non interacting, and non magnetized
976: hydrodynamic-like turbulent fluid. Such time scales are essential to
977: estimate spectral energy transfer (or cascades) rates ($\varepsilon
978: \simeq E(k)/\tau_{nl}$, where $E(k)$ is turbulent energy per unit
979: mode) in a hydrodynamic turbulent fluid. For a non-ideal, interacting
980: and magnetized fluid, incorporating complex interactions amongst the
981: turbulent eddies with a background or local mean magnetic field, the
982: nonlinear eddy interaction time needs to be modified. This was first
983: pointed out by Kraichnan (1965) in the context of magnetohydrodynamic
984: (MHD) fluids where the interaction of turbulent eddies with Alfv\'en
985: waves, excited as a result of a mean magnetic field, was
986: addressed. Owing to the presence of waves in a MHD fluid, turbulent
987: correlations between velocity and magnetic field and the corresponding
988: energy transfer time are determined primarily by $\tau \sim (b_0
989: k)^{-1}$, where $b_0$ is a typical amplitude of a local magnetic field
990: and is dimensionally identical to that of the velocity field by virtue
991: of Els$\ddot{\rm a}$sser's symmetry (Kraichnan 1965). Note carefully
992: the modification in the energy transfer (or cascade) time scale
993: because of the wave-turbulent eddy interaction process. Since the
994: convective time is the only nonlinear time scale involved in the
995: spectral transfer of turbulent energy, it it this time that accounts
996: entirely for the nonlinear cascade in the coupled plasma-neutral
997: heliosheath turbulence. Likewise, we introduce an analogous
998: modification in the coupled plasma-neutral ISM turbulence where the
999: energy cascade time scale due to charge exchange interactions amongst
1000: turbulent eddies, is determined explicitly by the factor $k_{ce}/k$
1001: (where the symbols have their usual meanings). It is this factor that
1002: determines how plasma and neutral fluids are coupled in heliosheath
1003: turbulence. Accordingly, the nonlinear spectral transfer time or the
1004: energy cascade rate is modified and is proportional to the factor
1005: $k_{ce}/k$ such that the new nonlinear eddy interaction time in the
1006: coupled plasma-neutral heliosheath turbulence is $\tau_{NL} \sim
1007: (k_{ce}/k) (1/kv_k)$. The interaction time $\tau_{NL}$ obtained in
1008: this manner for our coupled plasma-neutral problem not only adequately
1009: reproduces all three different regimes of interactions, but it also
1010: consistently describes the underlying physics in the context of
1011: heliosheath turbulence.
1012:
1013:
1014:
1015:
1016: On using the fact that $k_{ce}$ is typically larger than $k$ (or
1017: $k_c$, the characteristic turbulent mode, as defined elsewhere in the
1018: paper), i.e. $k_{ce}/k > 1$ in the ISM
1019: \citep{Florinski2003,Florinski2005,zank1999}, the new nonlinear time
1020: is $k_{ce}/k$ times bigger than the old nonlinear time i.e.
1021: $\tau_{NL} \sim (k_{ce}/k) \tau_{nl}$. This enhanced nonlinear
1022: interaction time in ISM turbulence is likely to prolong turbulent
1023: energy cascade rates. It is because of this enhanced or prolonged
1024: interaction time that a relatively large spectral transfer of
1025: turbulent modes tends to steepen the inertial range turbulent spectra
1026: in both plasma and neutral fluids. By extending the above
1027: phenomenological analysis, one can deduce exact (analytic) spectral
1028: indices of the inertial range decaying turbulent spectra, as
1029: follows. The new nonlinear interaction time-scale of ISM turbulence
1030: can be rearranged as
1031: \be
1032: \tau_{NL} \sim
1033: \frac{k_{ce}v_k}{kv_k}\frac{1}{kv_k}\sim \frac{\tau_{nl}^2}{\tau_{ce}},\ee
1034: where $\tau_{ce}\sim (k_{ce} v_k)^{-1}$ represents the charge exchange
1035: time scale. The energy dissipation rate associated with the coupled
1036: ISM plasma-neutral system can be determined from $\varepsilon \sim
1037: E_k/\tau_{NL}$, which leads to
1038: \be
1039: \varepsilon \sim \frac{v_k^2}{k_{ce}/k^2 v_k}\sim \frac{k^2 v_k^3}{k_{ce}}.\ee
1040: According to the Kolmogorov theory, the spectral cascades are local in
1041: $k$-space and the inertial range energy spectrum depends upon the
1042: energy dissipation rates and the characteristic turbulent modes, such
1043: that $E_k \sim \varepsilon^{\gamma} k^{\beta}$. Upon substitution of
1044: the above quantities and equating the power of identical bases, one
1045: obtains
1046: \be
1047: E_k \sim \varepsilon^{2/3} k^{-7/3}\ee
1048: for the plasma spectrum (the forward cascade inertial range). The spectral
1049: index associated with this spectrum, i.e. $7/3 \approx 2.33$, is
1050: consistent with the plasma spectrum observed in the (coupled ISM
1051: plasma-neutral) simulations (see \fig{fig5}{fig6}). Similar arguments in the
1052: context of neutral fluids, when coupled with the plasma fluid in ISM,
1053: lead to the energy dissipation rates
1054: \be
1055: \varepsilon \sim \frac{k^2 v_k^2}{k_{ce}/k^2 v_k}.\ee
1056: This further yields
1057: the forward cascade (neutral) energy spectrum
1058: \be
1059: E_k \sim \varepsilon^{2/3} k^{-11/3}\ee
1060: which is close to the simulation result shown in \Fig{fig6}.
1061:
1062:
1063: A key issue finally is to understand what role charge exchange modes
1064: play in coupled plasma-neutral ISM turbulence as they result
1065: essentially from the nonlinear charge exchange interactions. To
1066: address this issue, we plot charge exchange sources associated with
1067: the momentum and energy equations, \eqs{mhd}{hd}, in \Fig{fig7}. It
1068: appears from \Fig{fig7} that spectral energy is transferred
1069: predominantly at the larger scales by means of charge exchange mode
1070: coupling processes. The latter couples the large-scales, or smaller
1071: than $k_{ce}$ modes, efficiently to low-$k$ ISM turbulent modes. It
1072: is primarily because of this $k_{ce}-k$ mode-coupling in the smaller
1073: $k$ part of the spectrum of the coupled plasma-neutral ISM turbulence,
1074: that energy is pumped efficiently at the lower $k$ inertial range
1075: turbulent modes. The efficient coupling of the Fourier modes at low
1076: $k$'s further enhances the nonlinear eddy time-scales associated with
1077: the coupled plasma-neutral turbulence system which is consistent with
1078: the scaling $\tau_{NL} \sim (k_{ce}/k) \tau_{nl}$, where $k_{ce}/k
1079: >1$, as described above. This consequently leads to the steepening of
1080: the inertial range spectra observed in \fig{fig5}{fig6}. By contrast,
1081: higher $k$ modes, far from the energy cascade inertial range, are
1082: notably inefficient in transferring energy and momentum via charge
1083: exchange mode coupling interactions and are damped by small-scale
1084: dissipative processes in the coupled plasma-neutral ISM turbulence.
1085:
1086:
1087:
1088:
1089:
1090: \section{Conclusions}
1091: In conclusion, we have developed a self-consistent fluid model to
1092: describe nonlinear turbulent processes in a partially ionized and
1093: magnetized ISM gas. The charge exchange interactions couple the ISM
1094: plasma and the neutral fluids by exciting a characteristic charge
1095: exchange coupling mode $k_{ce}$, which is different from the
1096: characteristic turbulent mode $k$ of the coupled system. One of the
1097: most important points to emerge from our studies is that charge
1098: exchange modes modify the ISM turbulence cascades dramatically by
1099: enhancing nonlinear interaction time-scales on large scales. Thus on
1100: scales $\ell\ge\ell_{ce}$, the coupled plasma system evolves
1101: differently than the uncoupled system where large-scale turbulent
1102: fluctuations are strongly correlated with charge-exchange modes and
1103: they efficiently behave as driven (by charge exchange) energy
1104: containing modes of ISM turbulence. By contrast, small scale
1105: turbulent fluctuations are unaffected by charge exchange modes which
1106: evolve like the uncoupled system as the latter becomes less important
1107: near the larger $k$ part of the ISM turbulent spectrum (see also
1108: \Fig{fig7}). The neutral fluid, under the action of charge exchange,
1109: tends to enhance the cascade rates by isotropizing the ISM turbulence
1110: on a relatively long time scale. This tends to modify the
1111: characteristics of ISM turbulence which can be significantly different
1112: from the Kolmogorov phenomenology of fully developed turbulence. We
1113: believe that, it is because of this enhanced nonlinear eddy
1114: interaction time, that a large spectral transfer of turbulent energy
1115: tends to smear the current sheets in the magnetic field fluctuations
1116: and further cascade energy to the lower Fourier modes in the inertial
1117: range turbulent spectra. Consequently it leads to a steeper power
1118: spectrum. It is to be noted that the present model does not consider
1119: an external driving mechanism, hence the turbulence is freely
1120: decaying. Driven turbulence, such as due to large scale external
1121: forcing, e.g. supernova explosion, may force turbulence at larger
1122: scales. This can modify the cascade dynamics in a manner usually
1123: described by dual cascade process.
1124:
1125: There are issues that need further exploration. For example, neutrals
1126: are observed to evolve isotropically in our coupled plasma-neutral ISM
1127: turbulence model. A relevant question, whether their coupling to the
1128: plasma fluid at large-scales reduces spectral anisotropy in the plasma
1129: fluctuations, remains to be investigated. Furthermore, our model is
1130: currently 2D. It is then intriguing whether variations in the third
1131: dimension, i.e. 3D, introduce non trivial effects to our present 2D
1132: studies because turbulence in 2D differs considerably from that in 3D,
1133: particularly in the context of 3D neutral hydrodynamics fluid, where
1134: vortex stretching effect (which is absent in 2D) tears apart the
1135: large-scale structures and forbids the inverse cascade phenomena. By
1136: virtue of latter, 2D and 3D neutral fluids possess distinct spectral
1137: features characterized essentially by the number of the inviscid
1138: quadratic invariants \citep{biskamp}. How coupling of the plasma and
1139: neutral fluids in partially ionized ISM turbulence is modified by the
1140: 3D interactions remains a subject of our future investigations.
1141:
1142: We add finally that our self-consistent model can be useful in
1143: studying turbulent dynamics of partially ionized plasma in the
1144: magnetosphere of Saturn and Jupiter where outgassing from moons and Io
1145: and Encephalus introduces a neutral gas into the plasma.
1146:
1147:
1148:
1149:
1150:
1151:
1152:
1153: %\begin{figure}
1154: %\includegraphics{fig1.ps}% Here is how to import EPS art
1155: %\caption{\label{fig1} A figure caption. The figure captions are
1156: %automatically numbered.}
1157: %\end{figure}
1158:
1159: \acknowledgments
1160:
1161: %\begin{acknowledgments}
1162: The support of NASA(NNG-05GH38) and NSF (ATM-0317509) grants is acknowledged.
1163: %\end{acknowledgments}
1164:
1165:
1166:
1167:
1168:
1169:
1170:
1171:
1172: %% The reference list follows the main body and any appendices.
1173: %% Use LaTeX's thebibliography environment to mark up your reference list.
1174: %% Note \begin{thebibliography} is followed by an empty set of
1175: %% curly braces. If you forget this, LaTeX will generate the error
1176: %% "Perhaps a missing \item?".
1177: %%
1178: %% thebibliography produces citations in the text using \bibitem-\cite
1179: %% cross-referencing. Each reference is preceded by a
1180: %% \bibitem command that defines in curly braces the KEY that corresponds
1181: %% to the KEY in the \cite commands (see the first section above).
1182: %% Make sure that you provide a unique KEY for every \bibitem or else the
1183: %% paper will not LaTeX. The square brackets should contain
1184: %% the citation text that LaTeX will insert in
1185: %% place of the \cite commands.
1186:
1187: %% We have used macros to produce journal name abbreviations.
1188: %% AASTeX provides a number of these for the more frequently-cited journals.
1189: %% See the Author Guide for a list of them.
1190:
1191: %% Note that the style of the \bibitem labels (in []) is slightly
1192: %% different from previous examples. The natbib system solves a host
1193: %% of citation expression problems, but it is necessary to clearly
1194: %% delimit the year from the author name used in the citation.
1195: %% See the natbib documentation for more details and options.
1196:
1197:
1198:
1199:
1200: %% Use the figure environment and \plotone or \plottwo to include
1201: %% figures and captions in your electronic submission.
1202:
1203: %\clearpage
1204:
1205:
1206: \begin{thebibliography}{99}
1207:
1208: \bibitem[Banks (1966)]{banks}
1209: Banks, P., 1966,
1210: Planetary and Space Sci. 14, 1105.
1211:
1212: \bibitem[Basu et al (1994)]{basu}
1213: Basu, S., and Mouschovias, T. Ch. 1994, ApJ, 432, 720.
1214:
1215: \bibitem[Biskamp (2003)]{biskamp}
1216: Biskamp, D., 2003, Nonlinear MHD Turbulence, Academic Publisher.
1217:
1218: \bibitem[Cho \& Lazarian (2003)]{cho2003}
1219: Cho, J. and Lazarian, A., 2003, MNRAS, 345, 325.
1220:
1221: \bibitem[Elmegreen (1985)]{elmegreen}
1222: Elmegreen, B.G. 1985, ApJ, 299, 196.
1223:
1224: \bibitem[Fite et al (1962)]{fite}
1225: Fite, W. L., Smith, A. C. H., and Stebbings, R. F., 1962,
1226: Proc. R. Soc. London, em A. 268, 527.
1227:
1228: \bibitem[Florinski et al (2003)]{Florinski2003}
1229: Florinski, V., G. P. Zank, and N. V. Pogorelov (2003), Galactic cosmic
1230: ray transport in the global heliosphere, J. Geophys. Res., 108(A6),
1231: 1228, doi:10.1029/2002JA009695.
1232:
1233: \bibitem[Florinski et al (2005)]{Florinski2005}
1234: Florinski V., G. P. Zank, N. V. Pogorelov (2005), Heliopause stability
1235: in the presence of neutral atoms: Rayleigh-Taylor dispersion analysis
1236: and axisymmetric MHD simulations, J. Geophys. Res., 110, A07104,
1237: doi:10.1029/2004JA010879.
1238:
1239: \bibitem[Goldsmith \& Langer (1978)]{goldsmith}
1240: Goldsmith, P.F. and Langer, W.D. 1978, ApJ, 222, 881.
1241:
1242: \bibitem[Gayley et al (1996)]{gayley1996}
1243: Gayley, K., Zank, G. P., Pauls, H. L., Frisch, P. C., and Welty,
1244: D. E.: 1997, Astrophys. J. 487, 259.
1245:
1246:
1247: \bibitem[Groebner et al (2001)]{groebner2001}
1248: Groebner, R., Madhavi, M. A., Leonard, A. W., Osborne, T. H., Porter,
1249: G. D., Colchin, R. J., and Owen, L. W., 2001, GA Report GA-A23856,
1250: November; Groebner, R., Baker, D. R., and Burrell, K. H., et al, 2001,
1251: 41 No 12, 1789-1802.
1252:
1253:
1254: \bibitem[Hill \& Dessler (1990)]{hill1990}
1255: Hill, T. W.; Dessler, A. J., 1990
1256: Geophys. Res. Lett. (ISSN 0094-8276), vol. 17, Sept. 1990, p. 1677-1680.
1257:
1258:
1259:
1260: \bibitem[Heerikhuisen et al (2006)]{jacob2006}
1261: Heerikhuisen, J.; Florinski, V.; Zank, G. P. 2006,
1262: JGR., 111, A06110.
1263:
1264:
1265: \bibitem[Ip (1997)]{ip1997}
1266: Ip, W.-H., 1997,
1267: Icarus, Volume 126, Issue 1, pp. 42-57.
1268: DOI: 10.1006/icar.1996.5618
1269:
1270: \bibitem[Kolmogorov(1941)]{k41}
1271: Kolmogorov, A. N., 1941, Dokl. Akad. Nauk SSSR, 31, 538.
1272:
1273:
1274: \bibitem[Kraichnan (1965)]{kraichnan}
1275: Kraichnan, R. H., 1965, Phys. Fluids, 8, 1385.
1276:
1277: \bibitem[Kudoh et al (2007)]{kudoh}
1278: Kudoh, T., Basu, S., Ogata, Y., Yabe, T., 2007, MNRAS, 380, 499.
1279:
1280: \bibitem[Kulsrud \& Pearce (1969)]{kulsrud}
1281: Kulsrud, R. and Pearce, 1969, ApJ 156, 445.
1282:
1283:
1284: \bibitem[Leake et al (2005)]{leake}
1285: Leake, J. E., Arber, T. D., Khodachenko, M. L., 2005,
1286: arXiv:astro-ph/0510265v1.
1287:
1288: \bibitem[Li \& Nakamura (2004)]{li}
1289: Li, Z.-Y., Nakamura, F., 2004, ApJ, 609, L83.
1290:
1291: \bibitem[Muller et al (2006)]{muller2006}
1292: %Muller, H.-R. and Zank, G. P.,
1293: %2006, ASPC., 359, 276.
1294: Mueller, H.-R., Frisch, P. C., Florinski, V., and
1295: Zank, G. P., 2006, ApJ, 647, 1491.
1296:
1297: \bibitem[Mestel \& Spitzer (1956)]{mestel}
1298: Mestel, L., Spitzer, L., Jr., 1956, MNRAS, 116, 503.
1299:
1300: \bibitem[McNutt et al (1998)]{nutt}
1301: McNutt et al. 1998, JGR, 103, 1905.
1302:
1303: \bibitem[McIvor (1977)]{mcivor}
1304: McIvor, 1977, MNRAS 178, 85.
1305:
1306:
1307: \bibitem[Nakamura \& Li (2005)]{nakamura}
1308: Nakamura, F., Li, Z.-Y., 2005, ApJ, 631, 411.
1309:
1310:
1311: \bibitem[Nakano (1979)]{nakano}
1312: Nakano, T. 1979, PASJ, 31, 697.
1313:
1314: \bibitem[Oishi \& Mac Low (2006)]{oishi}
1315: Oishi, J. S. and Mac Low, M., 2006, ApJ, 638, 281.
1316:
1317: \bibitem[Padoan et al (2000)]{padoan}
1318: Padoan, P., Zweibel, E., and Nordlund, A., 2000, ApJ, 540, 332.
1319:
1320:
1321: \bibitem[Potekhin et al (2005)]{Potekhin2005}
1322: Potekhin, A. Y.; Lai, Dong; Chabrier, G.; Ho, W. C. G., 2005,
1323: Ad. Sp. Res., 35, 1158.
1324: DOI:10.1016/j.asr.2004.12.010.
1325:
1326: \bibitem[Pauls et. al (1995)]{pauls1995}
1327: Pauls, H. L., Zank, G. P., and Williams, L. L., 1995,
1328: J. Geophys. Res. A11, 21595.
1329:
1330:
1331: \bibitem[Scalo (1977)]{scalo}
1332: Scalo, J.M. 1977, ApJ, 213, 705.
1333:
1334: \bibitem[Shaikh et al (2006)]{shaikh2006}
1335: Shaikh, D. Pogorelov, N. V, and Zank, G. P., 2006,
1336: Astronomical Society of the Pacific
1337: Conference Series (ASPC), {\it Numerical
1338: Modeling of Space Plasma Flows: Astronum-2006}, Edited by N. V. Pogorelov and G. P. Zank,
1339: 359, 91.
1340:
1341: \bibitem[Shaikh \& Zank (2007)]{shaikh2007}
1342: Shaikh, D. and Zank, G. P., 2007, AIP Conf. Procs,
1343: 932, 111, DOI:10.1063/1.2778952.
1344:
1345: \bibitem[Shebalin \& Montgomery (1983)]{shebalin}
1346: Shebalin, J. V., Matthaeus, W. H., and Montgomery, D.
1347: 1983, J. Plasma Phys., 29, 525.
1348:
1349: \bibitem[Shu (1983)]{shu}
1350: Shu, F. H. 1983, ApJ, 273, 202.
1351:
1352: \bibitem[Singh et al (2004)]{singh2004}
1353: Singh, R., Rogister, A., and Kaw, P., 2004,
1354: Phys. Plasma, 11, 129.
1355:
1356: \bibitem[Slavin \& Frisch (2006)]{slavin2006}
1357: Slavin, J. D. and Frisch P. C., 2006,
1358: ApJ {\bf 651}, L37-L40.
1359:
1360: \bibitem[Valovie et al (2001)]{valovie2001}
1361: Valovie, M., Fielding, S. J., Carolan, P. G., et al., 2001,
1362: Contr. Fusion and Plasma Phys. 25A, 1829.
1363:
1364: \bibitem[Valovie et al (2002)]{valovie2002}
1365: Valovie, M., Carolan, P. G., Fielding, S. J., et al., 2002,
1366: Plasma Phys. Contr. Fusion 44, A175.
1367:
1368:
1369: \bibitem[Wood et. al (2000)]{wood2000}
1370: Wood, B. E.; Mueller, H.-R.; Zank, G. P, 2000,
1371: ApJ., 542, 493.;
1372: Wood, B. E.; Linsky, J. L.; Zank, G. P., 2000,
1373: ApJ., 537, 304.
1374:
1375: \bibitem[Zank \& Pauls (1996)]{zank1996}
1376: Zank, G. P. and Pauls, H. L., 1996, Sp. Sci. Rev., 78, 95.
1377:
1378: \bibitem[Zank et al. (1996)]{zank1996b}
1379: Zank, G. P.; Pauls, H. L.; Williams, L. L.; Hall, D. T., 2006,
1380: JGR, 101, A10, p. 21639-21656.
1381:
1382:
1383:
1384: \bibitem[Zank (1999)]{zank1999}
1385: Zank, G. P., Sp. Sci. Rev., 89, 413-688, 1999.
1386:
1387:
1388: \bibitem[Zank et al (2000)]{zank2000}
1389: Zank, G. P., Lu, J. Y., Rice, W. K. M., and Webb, G. M., 2000, J. Plasma
1390: Phys. {\bf 64}, 507-541.
1391:
1392: \bibitem[Zank et al (2006)]{zank2006}
1393: %Heliospheric Variation in Response to Changing Interstellar
1394: %Environments by
1395: Zank, G. P., Mueller, H. -R., Florinski, V., and
1396: Frisch, P. C., 2006, Chapter 2 in Solar Journey: Significance of our
1397: Galactic Environment for the Heliosphere and Earth, editor
1398: P.C. Frisch, Springer, pp. 281-316.
1399:
1400:
1401: \bibitem[Zweibel \& Josafatsson (1983)]{zweibel}
1402: Zweibel, E.G. and Josafatsson, K. 1983, ApJ, 270, 511.
1403:
1404:
1405: \end{thebibliography}
1406:
1407:
1408:
1409: \begin{figure}
1410: \plotone{fig1.ps} \figcaption{\label{fig0} Schematic illustrating
1411: the plasma neutral simulation regions both in the local interstellar
1412: medium and in the outer heliosheath (see \cite{zank1999} for a
1413: review of the solar wind interaction with the local ISM).}
1414: \end{figure}
1415:
1416: \clearpage
1417:
1418: \begin{figure}
1419: \plotone{fig2.ps} \figcaption{\label{fig1} Schematic
1420: illustrating the significance of the charge exchange mode, which
1421: corresponds to charge exchange interactions per unit area in the
1422: computational domain. This mode, being different from the
1423: characteristic turbulent mode $k$, is excited naturally when plasma
1424: and neutral fluids are coupled in the ISM by charge exchange. }
1425: \end{figure}
1426:
1427: \begin{figure}
1428: \plotone{fig3.ps} \figcaption{\label{fig3} ({\bf Left column}) Mode
1429: structures of plasma when coupled with the neutral ISM fluid. The
1430: local ISM case is considered where typical LISM parameters are;
1431: $\gamma=5/3, n_{0_{plasma}} = 0.06 cm^{-3}, n_{0_{neutral}} = 0.18
1432: cm^{-3}, T = 6500 K, V_{th} = 10 km/s, \sigma = 6.9\times 10^{-15},
1433: l_0 = 200 AU, L({\rm box length}) = 5 l_0$. ({\bf Right column}) The
1434: corresponding quantities for an uncoupled ISM plasma simulation
1435: measured at a similar time step (i.e. $t=20$ in normalized units).
1436: Notice the magnetic field structures in the uncoupled case showing
1437: the formation of thin (or sheet-like) structures unlike the coupled
1438: system. Also notice the differences in the density fields. }
1439: \end{figure}
1440:
1441: \clearpage
1442:
1443: \begin{figure}
1444: \plotone{fig4.ps} \figcaption{\label{fig4} The corresponding mode
1445: structures of neutral fluid in the presence of a plasma
1446: fluid. Simulation parameters are described in the previous figure.
1447: }
1448: \end{figure}
1449:
1450: \clearpage
1451:
1452: \begin{figure}
1453: \plotone{fig5.ps} \figcaption{\label{fig5}The inertial range
1454: turbulent spectrum for a neutral fluid shows a $k^{-3.2}$ spectrum when
1455: coupled through charge exchange to a plasma fluid. The spectral
1456: index exhibited by this spectrum is larger than that of an ordinary
1457: (hydrodynamic) fluid. The latter exhibits a Kolmogorov-like $k^{-3}$
1458: spectrum in the forward cascade regime (drawn schematically). The
1459: numerical resolution for our spectral studies is $1024^2$ modes in
1460: 2D.}
1461: \end{figure}
1462:
1463:
1464: \clearpage
1465:
1466: \begin{figure}
1467: \plotone{fig6.ps} \figcaption{\label{fig6} A similar steepening of
1468: turbulent spectra occurs in the plasma fluid when coupled to the
1469: neutrals through charge exchange. An uncoupled magnetofluid plasma
1470: exhibits a Kolmogorov-like $k^{-5/3}$ spectrum in the forward
1471: cascade regime (drawn schematically).}
1472: \end{figure}
1473:
1474:
1475: \clearpage
1476:
1477: \begin{figure}
1478: \plotone{fig7.ps} \figcaption{\label{fig7}
1479: Charge exchange spectrum associated with momentum and energy in both plasma
1480: and neutral fluids. Clearly the energy cascades are dominated by large-scale
1481: interactions, whereas small scales are not efficient enough to influence the
1482: turbulent spectra in the local ISM.
1483: }
1484: \end{figure}
1485:
1486: \clearpage
1487:
1488:
1489:
1490:
1491:
1492:
1493: \end{document}
1494:
1495: \be
1496: \label{mhd:cont}
1497: \frac{\partial \rho_p}{\partial t} + \nabla \cdot (\rho_p{\bf U}_p)=0;
1498: \ee
1499: \be
1500: \label{mhd:mom}
1501: \rho_p \left( \frac{\partial}{\partial t} + {\bf U}_p \cdot \nabla \right) {\bf U}_p
1502: = -\nabla P_p + \frac{1}{c} {\bf J} \times {\bf B}+{\bf Q}_M({\bf U}_p,{\bf V}_n, \rho_p, \rho_n, T_n, T_p);
1503: \ee
1504: \be
1505: \label{mhd:mag}
1506: \frac{\partial {\bf B}}{\partial t} = \nabla \times ({\bf U}_p \times {\bf B});
1507: \ee
1508: \eqa
1509: \label{mhd:en}
1510: \frac{\partial}{\partial t} \left( \frac{1}{2}\rho_p U_p^2 + \frac{P_p}{\gamma-1}+\frac{B^2}{8\pi}\right)
1511: &+& \nabla \cdot \left( \frac{1}{2}\rho_p U_p^2{\bf U}_p
1512: + \frac{\gamma}{\gamma-1}\frac{P_p}{\rho_p}\rho_p{\bf U}_p
1513: + \frac{c}{4\pi}{\bf E} \times {\bf B} \right) \nonumber \\
1514: &=&Q_E({\bf U}_p,{\bf V}_n,\rho_p, \rho_n, T_n, T_p).
1515: \eeq
1516:
1517:
1518:
1519: \be
1520: \label{hd:cont}
1521: \frac{\partial \rho_n}{\partial t} + \nabla \cdot (\rho_n{\bf V}_n)=0;
1522: \ee
1523: \be
1524: \label{hd:mom}
1525: \rho_n \left( \frac{\partial }{\partial t} + {\bf V}_n \cdot \nabla \right) {\bf V}_n
1526: = -\nabla P_n + {\bf Q}_M({\bf V}_n,{\bf U}_p,\rho_p, \rho_n, T_n, T_p);
1527: \ee
1528: \be
1529: \label{hd:en}
1530: \frac{\partial}{\partial t} \left( \frac{1}{2}\rho_n V_n^2 + \frac{P_n}{\gamma-1} \right)
1531: + \nabla \cdot \left( \frac{1}{2}\rho_n V_n^2{\bf V}_n
1532: + \frac{\gamma}{\gamma-1}\frac{P_n}{\rho_n}\rho_n{\bf V}_n \right)
1533: =Q_E({\bf V}_n,{\bf U}_p,\rho_p, \rho_n, T_n, T_p).
1534: \ee
1535:
1536:
1537:
1538: \bibitem[Zank et al (1998)]{zank1998}
1539: Zank, G. P.; Matthaeus, W. H.; Bieber, J. W.; Moraal, H., 1998,
1540: JGR., 103, 2085.
1541: