1: \documentclass[12pt,preprint]{aastex}
2: \bibpunct[, ]{(}{)}{;}{a}{,}{,}
3: %\documentstyle[12pt,aaspp4]{article}
4: \newcommand{\be}{\begin{equation}}
5: \newcommand{\ee}{\end{equation}}
6: \begin{document}
7: \title{Magnetic Field Structure and
8: Stochastic Reconnection in a Partially Ionized Gas}
9:
10:
11: \author{ A. Lazarian}
12:
13: \affil{Dept. of Astronomy, University of Wisconsin, 475 N Charter Street,
14: Madison WI 53706}
15:
16: \email{lazarian@astro.wisc.edu}
17:
18: \author{Ethan T. Vishniac}
19:
20: \affil{Department of Physics and Astronomy,
21: Johns Hopkins University,
22: Baltimore MD 21218}
23: \email{ethan@pha.jhu.edu}
24:
25: \and
26:
27: \author{ Jungyeon Cho}
28:
29: \affil{Dept. of Astronomy, University of Wisconsin, 475 N Charter Street,
30: Madison WI 53706}
31:
32:
33: \email{cho@astro.wisc.edu}
34:
35: \begin{abstract}
36: We consider stochastic reconnection in a magnetized, partially ionized
37: medium. Stochastic reconnection is a generic effect, due to field
38: line wandering, in which the speed of reconnection is determined by
39: the ability of ejected plasma to diffuse away from the current sheet
40: along magnetic field lines, rather than by the details of current
41: sheet structure. As in earlier work, in which we dealt with a fully
42: ionized plasma, we consider the limit of weak stochasticity, so that
43: the mean magnetic field energy density is greater than either the
44: turbulent kinetic energy density or the energy density associated with
45: the fluctuating component of the field. For specificity, we consider
46: field line stochasticity generated through a turbulent cascade, which
47: leads us to consider the effect of neutral drag on the turbulent
48: cascade of energy. In a collisionless plasma, neutral particle
49: viscosity and ion-neutral drag will damp mid-scale turbulent motions,
50: but the power spectrum of the magnetic perturbations extends below
51: the viscous cutoff scale. We give a simple physical picture of the
52: magnetic field structure below this cutoff, consistent with numerical
53: experiments. We provide arguments for the re\"emergence of the
54: turbulent cascade well below the viscous cut-off scale and derive
55: estimates for field line diffusion on all scales. We note
56: that this explains the persistence of a single power law form for the
57: turbulent power spectrum of the interstellar medium, from scales of
58: tens of parsecs down to thousands of kilometers. We find that under
59: typical conditions in the ISM stochastic reconnection speeds are
60: reduced by the presence of neutrals, but by no more than an order of
61: magnitude. However, neutral drag implies a steep dependence on the
62: Mach number of the turbulence. In the dense cores of $H_2$ regions
63: the reconnection speed is probably determined by
64: tearing mode instabilities.
65: \end{abstract}
66:
67: \keywords{Magnetic fields; Galaxies: magnetic fields,
68: ISM: molecular clouds, magnetic fields; Stars: formation }
69:
70:
71: \section{Introduction}
72:
73: One of the fundamental properties of astrophysical magnetic fields
74: is their ability to change topology via reconnection
75: \citep[see][]{PF00}. It is impossible to understand the
76: origin and evolution of large scale magnetic fields without
77: understanding the mobility of magnetic field lines. In a typical
78: astrophysical plasma, resistivity is very small
79: and flux freezing, which follows from assuming zero
80: resistivity, should be an excellent guide to the
81: motion of magnetic fields. In spite of this, fast
82: magnetic dynamo theory
83: \citep[see][]{P79,M78,KR80} invokes a
84: constantly changing magnetic field topology and motions during
85: which field lines cross each other\footnote{While standard mean-field
86: dynamo theory has been severely criticized on theoretical
87: and numerical grounds \citep{VC92,GD94,GD96,CH96,HCK96,B01},
88: an alternative version can be
89: formulated which evades these criticisms \citep[][]{VC01} and which
90: assumes fast reconnection only on two dimensional surfaces.}.
91: This assumption is supported
92: by observations of the solar magnetic field
93: \citep[see][and references contained therein]{D96,IIAW97}
94: which are difficult to explain unless flux
95: freezing is routinely violated on time
96: scales short compared to resistive time scales, at least within
97: thin current sheets.
98:
99: The Sweet-Parker model of reconnection is the simplest and most robust
100: \citep{P57, S58}. In this model, reconnection takes place within
101: a thin current sheet, which separates two large volumes containing
102: uniform, and very different, magnetic fields. The resulting reconnection speed
103: is less than the Alfv\'en speed
104: by the square root of the Lundquist number
105: $\sim {\cal R}_L^{-1/2}=(\eta/V_A L)^{1/2}$, where
106: $\eta$ is the resistivity, $V_A$ is the Alfv\'en speed, and $L$
107: is the length of the current sheet, assumed to be determined
108: by the large scale geometry of the problem. Under typical astrophysical
109: conditions this is very slow (e.g. for the Galaxy as a whole
110: ${\cal R}_L\sim 10^{20}$). This reconnection
111: speed is set by a geometrical constraint. Indeed,
112: plasma tied to the reconnecting magnetic field lines must be ejected
113: from the ends of the narrow current sheet. The disparity of scales, one
114: of which is macroscopic/astrophysical, while the other is microscopic,
115: i.e. determined by ohmic diffusion, makes the reconnection slow.
116:
117: This evident shortcoming of the Sweet-Parker reconnection
118: scheme has stimulated interest in
119: alternative models that allow fast reconnection. Although the
120: literature on magnetic reconnection is extensive
121: \citep[e.g.][and references therein]{PF00} it does not
122: successfully address this question. Models
123: that invoke an X-point reconnection geometry \citep{P64} have
124: been
125: shown to be unstable for sufficiently high $R_L$ \citep[see][]{B96},
126: while anomalous resistivity fails to provide rapid reconnection
127: under most astrophysical conditions \citep[see][]{P79}.
128: A general review of astrophysical magnetic reconnection
129: theory can be found in \citet{BMW03}.
130:
131: A notable exception to this discouraging state of affairs is
132: the recent work on fast collisionless
133: reconnection
134: \citep[see also the discussion by Bhattacharjee, Ma and Wang 2001]
135: {BSD97,SDD98,SD98}.
136: This
137: work indicates that under some circumstances a kind of standing
138: whistler mode can stabilize an X-point reconnection region. However,
139: these studies have not demonstrated the possibility of fast reconnection
140: for generic field geometries. They assume that there are no
141: bulk forces acting to produce a large scale current
142: sheet, and that the magnetized regions are convex, which minimizes
143: the energy required to spread the field lines. In addition,
144: this mechanism requires a collisionless
145: environment, where the electron mean free path is less than the current sheet
146: thickness \citep{TYJKC03, JYHK98}.
147: Unfortunately, in the laboratory the current sheet thickness
148: is comparable to the ion Larmor radius and
149: it is unclear how to generalize this criterion
150: to the interstellar medium, where the Sweet-Parker current sheet thickness
151: is typically much greater than the Larmor radius. {\it If} we require that
152: the ion skin depth, the characteristic scale of the standing Whistler mode,
153: be greater than the current sheet thickness, then we
154: have a criterion which is rarely satisfied in the interstellar medium.
155: In this paper we will concentrate on a mechanism that acts in turbulent media
156: and when it works,
157: produces rapid reconnection under a broad range of field geometries,
158: without regard to the particle collision rate. We will defer all discussion
159: of the relationship between stochastic reconnection and small scale
160: collisionless effects to a later paper.
161:
162: %In general, astrophysical magnetic reconnection takes place
163: %in turbulent media. \citet{PF00}
164: %have stressed that even if fast Petschek
165: %reconnection is possible it will still be necessary to
166: %demonstrate "that it will apply to the turbulent MHD regime''.
167: %If magnetic fields
168: %are turbulent then the boundary conditions for the current sheets
169: %change stochastically.
170: %On the other hand, boundary conditions need to be fine tuned for the
171: %Petschek reconnection scheme \citep[see][]{PF00}.
172:
173: In a previous paper \citep[henceforth LV99]{lv99}
174: we discussed `stochastic reconnection', a process which is
175: similar to Sweet-Parker reconnection, except that stochastic
176: wandering of the field lines produces a broad outflow region.
177: The properties of the outflow region are insensitive to the width of the
178: current sheet (and the value of ${\cal R}_L$), but depend on the
179: level of field line stochasticity. In a sufficiently noisy environment
180: the reconnection speed becomes a large fraction of the Alfv\'en speed.
181: In an {\it extremely} quiet environment the field lines do not
182: enter or leave the current sheet over its entire length and we
183: recover the Sweet-Parker reconnection model.
184: In LV99 we dealt with an inviscid and totally ionized fluid.
185: \footnote{In this paper we shall show that
186: this approximation is valid for partially ionized collisionless plasma
187: up to a certain percentage of neutrals, and for collisional gases with
188: resistivity larger than viscosity.}
189:
190: The notion that magnetic field stochasticity might affect
191: current sheet structures is not unprecedented. In earlier
192: work \citet{S70} showed that in collisionless plasmas the
193: electron collision time should be replaced with the time a
194: typical electron is retained in the current sheet. Also
195: \citet{JM84} proposed that current diffusivity should be modified
196: to include diffusion of electrons across the mean field due
197: to small scale stochasticity. These effects will usually be small
198: compared to effect of a broad outflow zone containing both
199: plasma and ejected shared magnetic flux. Moreover,
200: while both of these effects
201: will affect reconnection rates, they are not sufficient to
202: produce reconnection speeds comparable to the Alfv\'en speed
203: in most astrophysical environments.
204:
205: It is important to distinguish between stochastic reconnection,
206: as discussed in LV99, and the more conventional notion of
207: turbulent reconnection. The latter usually involves substituting a
208: turbulent diffusivity for the resistivity, and involves
209: a degree of small-scale mixing which is forbidden
210: on energetic grounds \citep[see \S 2 in][and the references
211: contained therein]{P92}. A common variation of this hypothesis
212: is that instabilities
213: in the current sheet will produce a hugely broadened current sheet
214: and a large effective resistivity within it \citep[for a review see][]
215: {B00}. On the
216: contrary, the former is largely a topological effect, using conventional
217: estimates of resistivity, and the only strong mixing associated
218: with it has to do with the polarization of field lines crossing
219: the current sheet. (That is, one expects sharp gradients, in the
220: current sheet, in the component of the magnetic lines perpendicular
221: to the current sheet.)
222: In this sense, stochastic reconnection belongs
223: to the class of models which try to explain fast reconnection by
224: appealing to a current sheet geometry which is `natural' in some
225: sense, but evades the limits set by the Sweet-Parker model. (The
226: bulk of the discussion in \citet{PF00} is centered on laminar
227: three dimensional field configurations which can lead to similarly
228: rapid reconnection speeds.)
229:
230: In a recent paper Kim \& Diamond (2001) addressed the problem of stochastic
231: reconnection by calculating the turbulent diffusion rate for magnetic flux
232: inside a current sheet. They obtained similar turbulent
233: diffusion rates for both two dimensional and reduced three
234: dimensional MHD. In both cases the presence of turbulence had a
235: negligible effect on the flux transport. The authors pointed out
236: that this would prevent the anomalous transport of magnetic flux within the current
237: sheet and concluded that both 2D and 3D stochastic reconnection proceed
238: at the Sweet-Parker rate even if individual small scale reconnection
239: events happen quickly. If true this would be not only rule out
240: the LV99 reconnection scheme, but also any other fast reconnection scheme.
241: In general astrophysical plasmas are turbulent and if the
242: enhancement of the local reconnection speed, e.g. due to collisionless
243: effects (see Drake et al. 2001), is irrelevant then reconnection must always
244: be slow.
245:
246: However turbulent diffusion rates within the current sheet are irrelevant for the
247: process of stochastic reconnection or, for broadly similar reasons, fast
248: collisionless reconnection.
249: The basic claim in LV99 is that realistic magnetic field topologies allow multiple
250: connections between the current sheet and the exterior environment, which would
251: persist even if the stochastic magnetic field lines were stationary ("frozen in
252: time") before reconnection. This leads to global outflow constraints which
253: are weak and do not depend on the properties of the current sheet. In particular,
254: the analysis in LV99 assumed that the current sheet thickness is
255: determined purely by ohmic dissipation and that turbulent diffusion of the magnetic
256: field is negligible inside, and outside, the current sheet. The major uncertainty
257: in this model is the behavior of the reconnected flux elements, which are nearly
258: perpendicular to the current sheet and must undergo multiple reconnections before
259: being ejected. We note also that models of collisionless reconnection also evade
260: the objection posed by Kim and Diamond topologically, that is, by stabilizing
261: an X-point reconnection topology, and opening up the rest of the current sheet.
262:
263: "Hyper-resistivity" \citep{S85,BH86,HB87} is a more subtle attempt to derive fast reconnection
264: from turbulence within the context of mean-field resistive MHD. The form of the parallel
265: electric field can be derived from magnetic helicity conservation. Integrating
266: by parts one obtains a term which looks like an effective resistivity
267: proportional to the magnetic helicity current. There are several assumptions
268: implicit in this derivation, but the most important problem is that by
269: adopting a mean-field approximation one is already assuming some sort of
270: small-scale smearing effect, equivalent to fast reconnection. \citet{S88}
271: partially circumvented this problem by examining the effect of tearing
272: mode instabilities within current sheets. However, the resulting reconnection
273: speed enhancement is roughly what one would expect based simply on the
274: broadening of the current sheets due to internal mixing. This effect
275: does not allow us to evade the constraints on the global
276: plasma flow that lead to slow reconnection speeds, a point which
277: has been demonstrated numerically \citep{ML85}
278: and analytically (LV99). Nevertheless, we show in \S 4 that this effect may be
279: important in the densest and coldest parts of the ISM.
280:
281: A partially ionized plasma fills a substantial volume within
282: our galaxy and the earlier stages of star formation take
283: place in a largely neutral medium. This motivates our study
284: of the effect of neutrals on reconnection.
285: The role of ion-neutral collisions is not trivial. On one hand, they
286: may truncate the turbulent cascade, reducing the small scale
287: stochasticity and decreasing the reconnection
288: speed. On the other hand, the ability of neutrals to diffuse perpendicular
289: to magnetic field lines allows for a broader particle outflow
290: and enhances reconnection rates.
291:
292: Reconnection in partially ionized gases has been already studied by
293: various authors \citep{NMA92,ZB97}. In a
294: recent study \citep[henceforth VL99]{VL99}
295: we studied the diffusion of neutrals away from the reconnection zone
296: assuming anti-parallel magnetic field lines \citep[see also][]{HZ03a}
297: The ambipolar reconnection rates obtained in VL99, although large
298: compared with the Sweet-Parker model, are insufficient either
299: for fast dynamo models or for the ejection of magnetic flux prior
300: to star formation. In fact, the increase in the reconnection speed
301: stemmed entirely from the
302: compression of ions in the current sheet, with the consequent enhancement of
303: both recombination and ohmic dissipation. This effect is small
304: unless the reconnecting magnetic field lines are almost exactly
305: anti-parallel \citep{VL99,
306: HZ03b}. Any dynamically significant shared field component
307: will prevent noticeable plasma compression in the current
308: sheet, and lead to speeds practically indistinguishable from the standard
309: Sweet-Parker result. Since generic reconnection regions will have
310: a shared field component of the same order as the reversing component,
311: the implication is that reconnection and ambipolar diffusion do not
312: change reconnection speed estimates significantly.
313:
314: None of this work on reconnection in partially ionized
315: plasmas includes the effects of
316: stochasticity. We expect that in the presence of turbulence,
317: reconnection rates will be substantially enhanced, as they are in
318: completely ionized plasmas. To generalize the concept of stochastic
319: reconnection to partially ionized plasmas we need a model for the
320: small scale stochasticity of a turbulent magnetic field in a partially
321: ionized plasma. In this paper we will begin by considering this
322: problem, and then apply our results to the reconnection speed.
323:
324: Following LV99, we consider reconnection in the presence of a weakly
325: stochastic
326: magnetic field. Except for the presence of noise, we imagine a
327: reconnection event exactly like a generic Sweet-Parker reconnection
328: event. Two volumes with average magnetic fields that are of comparable
329: strength, but differing directions are in contact over a surface of
330: length $L$. Due to the stochastic nature of the fields, field lines come
331: into contact over many small patches (see Fig.~1).
332: For each individual patch the Sweet-Parker reconnection model
333: should be applicable (or at least constitute a minimal reconnection speed).
334: The enhancement of reconnection rates follows from two effects.
335: First, since individual field lines wander out of the narrow current
336: sheet relatively easily, the longitudinal patch size is much smaller
337: than the overall size of the system.
338: This reduces the effective value of ${\cal R}_L$ and raises
339: the local reconnection speed. Second, whereas in the Sweet-Parker
340: scheme magnetic field lines reconnect sequentially, in the
341: presence of field line wandering field
342: lines many independent patches are brought into direct contact
343: and can reconnect simultaneously.
344: %EV
345: As a consequence, the rate of reconnection of the magnetic flux
346: is increased by a large factor, whose exact value depends on the level of
347: noise in the system.
348: %AL
349: %As a result, the upper limit on the
350: %reconnection speed imposed by the requirement that plasma escape from
351: %the current sheet along field lines is unrealistically large, larger
352: %than the Alfv\'en speed. The actual reconnection speed is determined
353: %by the width of the volume containing the field lines that cross the
354: %current sheet at some point, and approaches the Alfv\'en speed when the
355: %local turbulence is strong.
356:
357: In LV99 the truncation of the turbulent cascade was assumed to be
358: due to resistivity. Consequently, the smallest scale of field line
359: wandering decreases as resistivity decreases, and the number of
360: independent patches in contact within the reconnection zone increases.
361: {}From this we concluded that for an idealized inviscid
362: fluid the reconnection rate does not depend on
363: fluid resistivity. It does depend on the level
364: of magnetic field stochasticity. In the specific case where
365: the field line stochasticity is caused by a turbulent cascade,
366: it depends on the amplitude of the turbulence. However, there
367: is no necessary connection between turbulent motions and the
368: reconnection speed. In particular, the rate of turbulent transport
369: of mean magnetic flux is assumed to be negligible in this model,
370: %EV
371: and even the complete absence of turbulent diffusion in the
372: current sheet would not reduce the stochastic reconnection rate
373: \citep[cf.][]{KD01}.
374: %%AL
375: %In other words, the rate of turbulent transport does not present
376: %a bottleneck for stochastic reconnection (cf. Kim \& Diamond 2001).
377:
378: To quantify stochastic reconnection we have to use a particular
379: description of turbulence.
380: Motions in a magnetized medium can be expanded into incompressible (Alfv\'en)
381: and compressible (fast and slow) modes. There are theoretical arguments
382: \citep{gs95,LG01} \citep[see][for a review]{CLV03a}
383: suggesting that the nonlinear cascade of power for these
384: modes proceeds separately, although not entirely independently.
385: Simulations in \citet{CL02} support this idea
386: and show that the \citet[henceforth GS95]{gs95}
387: scaling is applicable to Alfvenic part of the MHD cascade.
388: For our purposes the scalings of slow and fast modes \citep{CL02}, \citep{CL03}
389: are less important since they are subjected to collisionless
390: damping\footnote{Whether or not those damped modes are important depends
391: on the process studied. For instance, \citet{YL02} show that
392: for scattering of cosmic rays the residual small amplitude
393: fast modes are much more efficient than the Alfv\'en modes. This,
394: however, is not true for the field wandering that we consider in this
395: paper.}
396:
397: Here we assume that the
398: GS95 model describes incompressible turbulence above the ambipolar
399: damping scale\footnote{We note, however,
400: that our qualitative conclusions for reconnection rates should be
401: valid for other models of MHD turbulence (see LV99) as long as they are in
402: rough agreement with observational constraints and numerical simulations.}.
403: To describe MHD turbulence below the scale of
404: viscous damping we present a new model of magnetic field structure in this
405: regime.
406: This model is in rough agreement with numerical simulations by
407: \citet[henceforth CLV02b]{CLV02b}.
408: %
409:
410:
411: %To describe turbulence above the ambipolar damping scale
412: %we adopt the Goldreich-Sridhar (1995, henceforth GS95)
413: %picture of the
414: %AL
415: %incompressible
416: %
417: %turbulent
418: %MHD cascade (see also the discussion in LV99). This picture is
419: %consistent with recent numerical simulations by
420: %\citet{CV00}, \citet{MG01}, and \citet{CLV02a}
421: %\footnote{We note, however,
422: %that our qualitative conclusions for reconnection rates should be
423: %valid for other models of MHD turbulence (see LV99) as long as they are in
424: %rough agreement with observational constraints and numerical simulations.}.
425: %To describe the turbulence below the scale of
426: %viscous damping below we present a new model of magnetic field structure.
427: %This model is in rough agreement with numerical simulations by
428: %\citet[henceforth CLV02b]{CLV02b}.
429:
430: In \S 2 of this paper we will consider the effect of a large
431: neutral fraction on a strongly turbulent cascade in a magnetized
432: plasma. In \S 3 we apply this to the problem of reconnection
433: in partially ionized plasmas. In \S 4 we apply this work to various
434: phases in the ISM. Finally, \S 5 contains our basic conclusions.
435:
436: \section{Magnetohydrodynamic Turbulence in a Partially Ionized Plasma}
437:
438: In this section we will consider the effect of neutral particles on the turbulent cascade
439: in the ISM. We begin by briefly reviewing the nature of the strong turbulent
440: MHD cascade and the dynamical influence of neutral particles. In \S 2.2 we
441: describe the cascade when viscous damping, due to neutral particles, is
442: strong, but the one fluid approximation remains valid. In \S 2.3 we consider
443: the uncoupled regime, covering scales where the neutral particles exert
444: a uniform drag on all motions. We end, in \S 2.4, with a brief discussion
445: of the implications of this picture for observations of turbulence in the ISM.
446:
447: \subsection{Neutral-ion damping}
448:
449: The role of neutral-ion damping in MHD turbulence has been discussed
450: previously in the context of the ISM \citep[in particular, see][]{S91,MS97}.
451: The basic conclusion was that neutral fluid heating is a plausible sink
452: for the turbulent energy revealed through measurements of interstellar
453: scintillation. Here we are concerned instead with how a neutral
454: gas component will modify the turbulent power spectrum. The most
455: relevant observational point is that the ISM turbulent power
456: spectrum has no strong features at wavelengths where neutral-ion
457: coupling would be expected to play a dominant role \citep{ARS95}.
458: Instead, the power spectrum extends to very small scales ($<10^8$ cm)
459: in an approximate power law.
460: Qualitatively, this suggests that stochastic reconnection can
461: take place even in partially neutral plasmas. However, several
462: basic questions remain unanswered.
463: Previous work on turbulence in the ISM has not included
464: a discussion of the most plausible model for MHD turbulence (although
465: a simple hydrodynamic model was addressed, which is remarkably
466: close to the model we use here). Also, we need to understand
467: why neutral damping fails to produce a strong signature in the
468: ISM, or at least in the diffuse ionized component of the ISM, before
469: we can construct a general model for its role in partially ionized
470: plasmas.
471:
472:
473: \subsubsection{The Goldreich-Sridhar model}
474:
475: The GS95 model of strong MHD turbulence is based on the notion of a Kolmogorov-like
476: cascade with an anisotropy imposed by the large scale magnetic field.
477: The exact degree of anisotropy follows from an average balance between
478: hydrodynamic and magnetic forces.
479: Eddies on a given scale are characterized by a wavenumber perpendicular
480: to the mean magnetic field direction, $k_{\perp}$, and a parallel
481: wavenumber, $k_{\|}$, such that the rate of eddy turnover time is
482: equal to the rate wave propagation along magnetic field, i.e.
483: \be
484: k_{\perp}v_k\approx k_{\|}V_A~~~,
485: \label{balance}
486: \ee
487: where $v_k$ is the typical velocity at the scale characterized
488: by the wavenumber $(k_{\|},k_{\perp})$. As $v_k$ goes down with the
489: increase of $k_{\bot}$ this condition implies
490: eddies that are elongated along the field direction, and become
491: more elongated as we go to smaller scales. In order to simplify our
492: notation, we will refer to $k_{\perp}$ below as $k$.
493:
494: If energy is injected isotropically on some scale $l$, with $v_l\le V_A$, then
495: the cascade will begin in a regime of weak turbulence, in which motions can
496: be characterized as weakly interacting waves, with a frequency $\omega=k_{\|}V_A\sim $constant and
497: a nonlinear decay rate (see discussion in a review by Cho, Lazarian \&
498: Vishniac 2003)
499: \be
500: \tau_{nl}^{-1} \sim {k^2v_k^2l\over V_A},
501: \ee
502: so that conservation of energy implies
503: \be
504: v_k\sim v_l(kl)^{-1/2}.
505: \ee
506: For $\tau_{nl}^{-1}$ less than the wave frequency
507: the corresponding scale-dependent diffusion coefficient will be (LV99)
508: \be
509: D_k\sim {v_k^2\over\omega^2\tau_{nl}}\sim {v_l^4l\over V_A^3},
510: \ee
511: which is actually independent of scale. The decrease in wave motions at
512: larger $k$ is
513: balanced by the decrease in the coherence time of the waves.
514: Here we will ignore the contribution of these scales to turbulent
515: diffusivity in
516: favor of the contribution from smaller, strongly turbulent, scales.
517:
518: The weak turbulent cascade ends at a scale $k_T$ where equation (\ref{balance})
519: is satisfied. At this scale
520: \be
521: v_T\sim {v_l^2\over V_A}, \hbox{\ and\ } k_Tl\sim \left({V_A\over v_l}\right)^2.
522: \ee
523: At larger $k$ the strong turbulence model applies and (see GS95)
524: \be
525: k_{\|} \approx l^{-1} \left({k\over k_T}\right)^{2/3}.
526: \label{k_p2m}
527: \ee
528: The rate of turbulent energy transfer is $k_{\|}V_A$, which means
529: \be
530: \tau_{nl}^{-1}\approx {V_A\over l}\left({k\over k_T}\right)^{2/3},
531: \label{tau2m}
532: \ee
533: while the rms turbulent fluid velocity is given by
534: \be
535: v_k\approx v_T \left({k\over k_T}\right)^{-1/3}.
536: \label{velm}
537: \ee
538: The magnetic field perturbations, $b_k$, are
539: \be
540: b_k\approx v_k (4\pi\rho)^{1/2}.
541: \label{belm}
542: \ee
543:
544: This model presupposes that the turbulent velocities are subalfvenic,
545: and we adopt this assumption in the rest of this paper. This is
546: less restrictive than it might appear, since as long as there is
547: some scale $l'$ in the turbulent cascade where $v_{l'}\sim V_A$ we
548: can take $l=l'\approx k_T^{-1}$, $v_T=v_l=V_A$ and use this model of turbulence
549: for all smaller scales.
550: %AL
551: Moreover, if the turbulent energy is larger than the magnetic
552: energy, we can expect rapid growth of the magnetic field through the
553: turbulent dynamo, that is, the generation of a disordered field in
554: rough equipartition with the turbulent kinetic energy (see, however,
555: Schekochihin et al. 2003).
556:
557: The Goldreich-Sridhar scalings can be easily understood. They reflect
558: the fact that on small scales it is difficult to bend magnetic field lines,
559: but it is still easy to mix them up.
560:
561: In a fully ionized astrophysical plasma, shear viscosity
562: is generally less important than resistivity
563: in damping MHD turbulence. In a partially neutral medium a combination
564: of neutral particle viscosity and ion-neutral collisional coupling
565: drives damping. Following
566: GS95 we concentrate on the diffusion of momentum across field lines.
567:
568: \subsubsection{Ion-neutral decoupling: theoretical considerations}
569:
570: The preceding discussion assumes that the plasma is a single, tightly coupled,
571: fluid with negligible viscosity. Many astrophysical fluids are partially ionized, the
572: neutrals are imperfectly coupled to the ions. An obvious consequence is a substantially
573: increased viscosity since the neutrals can cross magnetic field lines.
574:
575: The coupling between ions and neutrals is determined by
576: the rate of ion-neutral collisions, which is
577: \be
578: t_{in}^{-1}={m_n\over m_n+m_i} n_n\langle v_r \sigma_{in}\rangle,
579: \ee
580: where $v_r$ is the ion-neutral relative velocity, $\sigma_{in}$ is
581: the ion-neutral collisional cross section, $m_i$ and $m_n$ are the
582: typical ion and neutral masses, $n_n$ is the neutral number density,
583: and angular brackets denote averaging. From \citep{DRD83}
584: we adopt ${\langle v_r \sigma_{in}\rangle}\approx 1.9\times 10^{-9}$
585: cm$^3$ s$^{-1}$. The rate at which neutrals exchange momentum with
586: ions is $t_{ni}^{-1}= t_{in}^{-1}\rho_i/\rho_n$.
587:
588: If the mean free path for a neutral particle, $l_n$,
589: in a partially ionized gas
590: with density $n_{tot}=n_n+n_i$ is much less than
591: the size of the eddies under
592: consideration, i.e. $l_n k\ll 1$, the damping time is
593: \be
594: t_{damp}\sim
595: \nu_n^{-1} k^{-2}\sim \left(\frac{n_{tot}}{n_n}\right)(l_n c_n)^{-1} k^{-2}~~~,
596: \label{tdamp}
597: \ee
598: where $\nu_n$ is the effective viscosity produced by neutrals and
599: $c_n$ is the sound speed in the neutrals
600: \footnote{The viscosity
601: across magnetic field lines, due to ion-ion collisions,
602: is typically small
603: as ion motions are constrained by the magnetic field. For
604: a collision rate much smaller than the ion cyclotron frequency
605: the ratio of ion perpendicular viscosity to resistivity is a few times
606: the ratio of the ion thermal pressure to the magnetic pressure.
607: In the collisional limit this is multiplied by a factor
608: of $(\Omega_i/\nu_i)^2\ll 1$, and resistivity efficiently
609: dissipates magnetic field perturbations on scales greater
610: than the viscous damping scale.
611: The drag coefficient for neutral-neutral collisions is
612: $\sim 1.5\times10^{-10} T^{1/3}$ cm$^3$ s$^{-1}$
613: with $T$ measured in Kelvins
614: \citep{S78},
615: so collisions with other neutrals will dominate
616: for $n_i/n_n$ less than $\sim 0.08 T^{1/3}$.}.
617:
618: Consider first a mostly neutral gas (i.e. $n_i\gg n_n$).
619: Turbulence can cascade to small scales if the
620: turbulent eddy rate $\tau^{-1}\sim k v_k$ is larger than
621: the viscous damping rate $t_{damping}^{-1}$. In a partially ionized
622: gas the one fluid approximation is valid if neutrals experience multiple collisions
623: with ions in an eddy turnover time, i.e. $t_{ni}^{-1}>\tau^{-1}$.
624: Therefore, MHD turbulence will exhibit Goldreich-Sridhar scaling
625: up to the damping scale if $t_{ni}^{-1}>t_{damping}^{-1}$. On
626: the other hand, if
627: $t_{ni}^{-1}<t_{damping}^{-1}$ neutrals will not follow the ions and
628: magnetic fields. Instead they will form a {\it hydrodynamic} cascade
629: as soon as the neutral-ion collisional rate
630: $t_{ni}^{-1}$ is of the order of the eddy turnover time
631: $\tau$. At this scale, and all smaller scales, the ionic fluid motions
632: will damp at
633: the rate of ion-neutral collisions $t_{in}^{-1}$, which
634: in mostly neutral gas is much larger than $t_{ni}^{-1}$.
635:
636: What will happen to the magnetic fields when ionic motions are damped?
637: Collisions between the ions and the neutral particles will prevent magnetic
638: tension from straightening the field lines efficiently.
639: Instead they will be moved,
640: entangled and stretched by large undamped eddies. As the eddy turnover
641: rate increases with the decrease of the scale, the marginally damped
642: eddies at the damping scale will the most important.
643: This implies a picture very different from the
644: Goldreich-Sridhar cascade.
645:
646: If neutrals constitute a tiny impurity in the ionized plasma,
647: it is clear that they cannot affect the MHD cascade. We shal
648: quantify this intuitive picture below.
649:
650: \subsubsection{Ion-neutral decoupling: a simple model}
651:
652: Consider first a toy model of ion-neutral interaction.
653: If we ignore viscous damping, the equations for the
654: ions and neutrals are:
655: \be
656: {v_i\over\tau} ={v_n-v_i\over t_{in}} -{\rho\over\rho_i}
657: \omega_A^2\tau v_i+F_i~~~,
658: \label{memovi}
659: \ee
660: and
661: \be
662: {v_n\over\tau} ={\rho_i\over\rho_n}{v_i-v_n\over t_{in}}+F_n~~~.
663: \label{memovn}
664: \ee
665: For simplicity we have used $F_i$ and $F_n$ to denote both pressure
666: forces and the nonlinear turbulent accelerations for the ions and neutrals,
667: respectively.
668:
669: To linear order equations, and ignoring sound waves, (\ref{memovi}) and (\ref{memovn}) give
670: the dispersion relation for Alfv\'en waves in a partially
671: ionized plasma \citep{M77}:
672: \be
673: {t_{in}\over\tau}\left(1+(\omega_A\tau)^2{\rho\over\rho_i}\right)
674: ={-1\over 1+\aleph}~~~,
675: \label{ldis}
676: \ee
677: where $\aleph$ is a dynamical coupling parameter, defined by
678: \be
679: \aleph\equiv {\rho_i\over\rho_n}{\tau\over t_{in}}~~~.
680: \ee
681: When the coupling is very tight, $\aleph\gg 1$, and we have the usual relation
682: for Alfv\'en and pseudo-Alfv\'en modes in a single fluid,
683: \be
684: \tau^{-2}=-\omega_A^2~~~,
685: \label{alf}
686: \ee
687: where $\tau$ is imaginary.
688:
689: As the turbulence cascades to smaller scales, $\tau$, and
690: consequently, $\aleph$, decreases.
691: If we express the value of $\aleph$ at the viscous damping scale
692: as $\aleph_c$, then we have two obvious alternatives.
693: Either $\aleph_c>1$ and two-fluid effects are negligible right up to
694: the damping scale, or $\aleph_c<1$ and the ions and neutrals decouple
695: in the middle of the turbulent cascade. In order to calculate
696: $\aleph_c$ we need to calculate viscous damping in the large $\aleph$
697: limit.
698:
699:
700: Combining Eqs~(\ref{k_p2m})
701: and
702: (\ref{tdamp}) we get
703: \be
704: {t_{damp}\over\tau}\sim f_n^{-1} \left(\frac{v_l}{V_A}\right)^{1/3}
705: \left(\frac{l_n}{l}\right)^{1/3} \left(\frac{v_l}{c_n}\right)
706: (l_n k)^{-4/3}~~~,
707: %\label{1damp}
708: \label{eq.4}
709: \ee
710: where $c_n$ is the sound speed and $f_n$ is the neutral fraction.
711: For most of our applications we will have $f_n\sim 1$.
712: The damping scale, $k_c^{-1}$, is defined by $t_{damp}\sim \tau_k$, so that
713: \be
714: k_c\sim l_n^{-1}\left(\frac{v_l}{c_n}\right)^{3/4}
715: \left(\frac{v_l}{V_A}\right)^{1/4}\left(\frac{l_n}{l}\right)^{1/4}
716: f_n^{-3/4},
717: \label{4.2.1}
718: \ee
719: \be
720: \tau_c^{-1}\sim k_c^2f_nc_nl_n\sim
721: \left(\frac{c_n}{l_n}\right)\left(\frac{v_l}{c_n}\right)^{3/2}
722: \left(\frac{l_n}{l}\right)^{1/2} \left(\frac{v_l}{V_A}\right)^{1/2}
723: f_n^{-1/2},
724: \label{4.2.2}
725: \ee
726: \be
727: k_{\|,c}\sim \tau^{-1}_s V_A^{-1}\sim l_n^{-1}
728: \left(\frac{v_l}{c_n}\right)^{1/2}\left(\frac{l_n}{l}\right)^{1/2}
729: \left(\frac{v_l}{V_A}\right)^{3/2}f_n^{-1/2},
730: \label{4.2.3}
731: \ee
732: and
733: \be
734: v_c\sim v_l
735: \left(\frac{l_n}{l}\right)^{1/4}
736: \left(\frac{c_n}{V_A}\right)^{1/4}f_n^{1/4},
737: \label{4.2.4}
738: \ee
739:
740: {}From the definition of
741: %$\aleph_c$
742: %AL
743: $\aleph_c\equiv \aleph (\tau_c)$
744: we see
745: that
746: \be
747: \aleph_c\sim f_n^{1/2}
748: \left({l\over l_n}\right)^{1/2} \left({\rho_i l_n\over \rho_n t_{in}c_n}\right)
749: \left({c_n\over v_l}\right)^{3/2}\left({V_A\over v_l}\right)^{1/2}.
750: \label{als}
751: \ee
752:
753: Equation (\ref{als}) seems to imply that $\aleph_c$ must always be greater than
754: one, but a closer examination suggests that for $\rho_i\ll \rho_n$ the third
755: term on the right hand side can be small enough to offset the second
756: term. In fact, the value of $\aleph_c$ has to be determined for each
757: situation. When $\aleph_c\ll 1$ equations (\ref{4.2.1}) through (\ref{4.2.4})
758: have no physical meaning, since
759: the plasma stops behaving as a single fluid when $\aleph$ drops below
760: one. Nevertheless, we can still use these expressions as useful parameterizations
761: of the turbulent cascade. In particular, if $\aleph_c<1$ then the scale of
762: decoupling, when $\aleph=1$, is given by
763: \be
764: k'\sim k_c\aleph_c^{3/2},
765: \label{decoup1}
766: \ee
767: while at higher wavenumbers equation (\ref{ldis}) becomes
768: \be
769: {t_{in}\over\tau}\left(1+\omega_A^2\tau^2{\rho\over\rho_i}\right)=-1~~~.
770: \label{memo6}
771: \ee
772: This expression has two roots. When the collision rate is so
773: small that the ions
774: and neutrals are completely decoupled we have usual
775: dispersion relation for Alfv\'en waves
776: \be
777: \tau^{-2}\sim -\omega_A^2 \rho/\rho_i~~~.
778: \label{decouple1}
779: \ee
780: When the collision rate is large the two roots are
781: \be
782: \tau^{-1}\sim -\omega_A^2t_{in}{\rho\over\rho_i},~~~ -{1\over t_{in}}.
783: \label{memo7}
784: \ee
785: The latter root corresponds to the case where magnetic forces are
786: negligible and the former is the usual ambipolar diffusion rate, when the magnetic
787: field pushes the ions through a neutral background. Neither limit
788: is appropriate for hydrodynamic turbulence in the neutral fluid,
789: in which case $v_n$, $\tau$, and ${\bf k}$ are imposed by
790: the turbulent cascade. From equation (\ref{memovn}) we see that
791: if the ions are prevented from moving by the magnetic field then
792: we get a damping rate $\sim ({\rho_i/\rho_n})t_{in}^{-1}$, which
793: will be negligible once the hydrodynamic cascade has proceeded
794: to eddies with turn over rates larger than the decoupling rate.
795: Since the hydrodynamic eddies will be approximately isotropic, equation
796: (\ref{memovi}) guarantees that $v_i\ll v_n$ for all scales below
797: the decoupling scale.
798:
799: The limit defined by equation (\ref{decouple1})
800: has been previously considered \citep[see, for example][]{KP69}.
801: We will finish this subsection by considering damping in this regime, and
802: the minimum neutral fraction required for neutral damping on any scale.
803:
804: It is easy to see that equation (\ref{decouple1}) corresponds to the regime
805: $l_n k\gg 1$ and the waves are damped
806: at a rate of $t_{in}^{-1}$. Therefore,
807: \be
808: \tau_k^{-1}t_{damp}\sim
809: \left(\frac{v_l}{c_n}\right)
810: \left(\frac{l_n}{l}\right)^{1/3}
811: \left(\frac{v_l}{V_A}\right)^{1/3}
812: \left({c_nt_{in}\over l_n}\right)
813: (l_n k)^{2/3}.
814: \label{o1}
815: \ee
816: If we have a turbulent cascade in the ions, with $\tau_{nl}^{-1}t_{in}>1$, then
817: the cascade proceeds to very small scales, and neutral damping plays no role
818: in its dynamics.
819: Although the ions are decoupled from the neutrals, we use the
820: usual value of $V_A$ in this expression since it appears only
821: in the combination $v_l^4/(lV_A)$, which is the energy cascade
822: per unit mass.
823: The ratio $c_nt_{in}/l_n$ is roughly
824: \be
825: {c_nt_{in}\over l_n}={l_{in}\over l_{nn}}+
826: {c_it_{in}\over c_nt_{ni}}
827: \approx {l_{in}\over l_{nn}}+
828: {\rho_i\over\rho_n}\left(\frac{m_n}{m_i}\right)^{1/2}.
829: \ee
830: If the ions and neutral particles have the
831: same mass, this will be $\sim f_n^{-1}$ when the neutral
832: fraction is small, and a number of order unity otherwise.
833:
834: Here we have assumed that the rms velocities of the ions and neutrals
835: are dominated by the thermal distribution.
836: Our treatment should still be
837: applicable when the turbulence is supersonic, but sub-Alfv\'enic
838: (see
839: discussion of the point in Cho, Lazarian \& Vishniac 2002a),
840: as long as we replace $c_i$ and $c_n$ with the appropriate
841: rms turbulent velocities. This
842: is most likely to happen on large scales, so that $l_n k\ll 1$.
843: In this case the viscosity contributed by neutrals stays
844: the same so that equation (\ref{eq.4}) does not
845: require any modifications. In the opposite limit, when $kl_n\gg 1$,
846: for supersonic turbulence $\langle v_i^2\rangle^{1/2}\approx V_A^{\star}$,
847: where $V_A^{\star}$
848: is the Alfv\'en velocity when ions and neutrals are decoupled, i.e.
849: $V_A^{\star}=(n_n m_n +n_i m_i)^{1/2}/(n_i m_i)^{1/2} V_A$.
850: However an increase in the ion velocity will result in a proportional
851: increase in the drag coefficient, and a consequent drop in $t_{in}$.
852: Thus the condition provided by equation (\ref{o1}) does not change.
853:
854: Equations (\ref{eq.4}) and (\ref{o1}) can be used to find the
855: minimal neutral fraction required to damp the turbulent cascade
856: at scales close to $l_n$. Together they imply a maximum damping
857: rate for $kl_n\sim 1$. We see $\tau_k^{-1}t_{damp}\sim 1$
858: at that scale requires a small neutral fraction, so that
859: $l_n\sim v_nt_{ni}$. Assuming a single thermal velocity we get
860: \be
861: f_n>f_{crit}=7\times 10^{-2}\left (\frac{l}{pc}\right)^{-1/3}
862: \left({v_T\over c_n}\right)^{2/3}\left({V_A\over c_n}\right)^{1/3}
863: n_{tot}^{-1/3}T^{1/6},
864: \label{fcrit}
865: \ee
866: as the minimal condition for dissipating an MHD turbulent cascade
867: through neutral particle viscosity.
868: Typically, turbulent motions
869: in the ISM are of order $c_n$ or greater, so only a small neutral
870: fraction is necessary to affect the turbulent cascade.
871: If there is a source of noise on scales
872: $\ll l_n$, such that damping given in equation (\ref{o1}) is ineffective,
873: then neutral friction can be ignored on all smaller scales.
874: However, we will show in the following pages that small scale turbulence
875: can appear in a partially neutral plasma even
876: in the absence of any small scale driving.
877:
878:
879: \subsection{The coupled regime: viscosity-damped turbulence}
880:
881: In this subsection we consider strong MHD turbulence with strong
882: neutral particle damping, but on scales where the one-fluid
883: approximation remains valid.
884:
885: In hydrodynamic turbulence viscosity sets a minimal scale for
886: motion, with an exponential suppression of motion on smaller
887: scales. Below the viscous cutoff the kinetic energy contained in a
888: wavenumber band is
889: dissipated at that scale, instead of being transferred to smaller scales.
890: This means the end of the hydrodynamic cascade, but in MHD turbulence
891: this is not the end of magnetic structure
892: evolution.
893: %\footnote{Qualitatively this may be understood in the following
894: %way. Mixing motions perpendicular to the direction of the magnetic field
895: %dominate the dynamics of the Alfvenic cascade and tangle magnetic field
896: %lines. The created structures cannot miraculously straighten up at the
897: %viscous damping scale. Instead, they keep evolving due to the
898: %shear associated with the smallest undamped eddies.}.
899: For viscosity much larger than resistivity,
900: $\nu\gg\eta$, there will be a broad range of
901: scales where viscosity is important but resistivity is not. On these
902: scales magnetic field structures will be created through a
903: combination of large scale shear and the small scale motions generated
904: by magnetic tension. As a result, we expect
905: a power-law tail in the energy distribution, rather than an exponential
906: cutoff.
907:
908: Here we discuss a conservative model for the damped regime.
909: It is motivated by simulations by \citet[henceforth CLV02b]{CLV02b}
910: and is consistent with those simulations \citep[see also][]{CLV03b}.
911: A complete
912: understanding of this regime will be deferred until
913: higher numerical resolution runs are available and a more detailed
914: theoretical treatment has been completed. Our goal here is a model
915: which can serve as an approximate guide. We will construct this
916: model using the notion of local interactions in phase space
917: \citep[cf.][]{Sp99}, a
918: constant cascade of energy to small scales, and a force balance
919: between magnetic
920: tension and viscous forces. We add to this two ingredients
921: that are directly suggested by the simulations: a constant curvature
922: for the field lines and an intermittent magnetic field distribution.
923:
924: To begin with we define a filling factor $\phi_k$, which is the
925: fraction of the volume containing strong magnetic field perturbations
926: with a scale $k^{-1}$. We denote the velocity and perturbed
927: magnetic field inside these subvolumes with a ``$\hat{\ } $'' so
928: that
929: \be
930: v_k^2=\phi_k \hat v_k^2,
931: \ee
932: and
933: \be
934: b_k^2=\phi_k \hat b_k^2.
935: \ee
936:
937: We note that at the critical damping scale, $k_c$ and
938: $k_{\|, c}$, the time-scale for inviscid turbulence is
939: \be
940: \tau^{-1}_s\approx k_c^2 \nu\approx k_{\|, c} V_A\approx k_c v_c
941: \approx k_c{b_c\over (4\pi\rho)^{1/2}}~~~.
942: \label{4.1}
943: \ee
944: Assuming we are in the coupled regime, we have $\aleph_c=\rho_i \tau_c/
945: (\rho_n t_{in})>1$. Motions on this scale are marginally damped, so that
946: smaller scale structures will be continuously sheared at a rate $\tau_c^{-1}$.
947: These structures will reach a dynamic equilibrium if they generate a
948: comparable shear, that is
949: \be
950: k \hat v_k\sim \tau_c^{-1}.
951: \label{s1}
952: \ee
953: The magnetic energy will cascade to higher wavenumbers at this
954: same rate, so that
955: \be
956: b_k^2k\hat v_k\sim {b_c^2\over \tau_c}.
957: \label{s2}
958: \ee
959: Consequently, $b_k\sim b_c$ for $k>k_c$.
960:
961: Next we assume that the curvature of the magnetic field lines
962: changes slowly, if at all, in the cascade. This is consistent
963: with a picture in which the cascade is driven by repeated shearing
964: at the same scale. It is also consistent with the numerical
965: work described in CLV02b, which yielded a constant $k_{\|}$
966: throughout the viscously damped nonlinear cascade.
967:
968: Finally, we can balance viscous and magnetic tension forces
969: to find
970: \be
971: k^2\nu \hat v_k \sim k\nu\tau_c^{-1}\sim \max[\hat b_kk_c,B_0k_{\|,c}] \hat b_k
972: \sim k_c\hat b_k^2.
973: \label{s3}
974: \ee
975: This suggests a picture in which the cascade to larger wavenumbers consists
976: of an evolution to increasing gradients perpendicular to both the
977: mean field direction and the local perturbed field component. This
978: implies rapidly increasing magnetic pressure gradients, but these
979: can compensated by plasma density fluctuations.
980:
981: Combining our results we find
982: \be
983: \hat b_k \sim b_c \left({k\over k_c}\right)^{1/2}, b_k\sim b_c,
984: \label{4.5}
985: \ee
986: \be
987: \hat v_k\sim v_c \left({k_c\over k}\right), v_k\sim
988: v_c\left({k_c\over k}\right)^{3/2},
989: \label{4.6}
990: \ee
991: and
992: \be
993: \phi_k\sim {k_c\over k}.
994: \label{4.6a}
995: \ee
996:
997: The associated velocities fall rapidly, although not exponentially.
998: In terms of one dimensional spectra we have
999: a magnetic energy spectrum $E^b(k)\sim k^{-1}$,
1000: and a kinetic energy spectrum $E^v(k)\sim k^{-4}$. These scaling
1001: laws should be compared
1002: with $E^v\sim E^k\sim k^{-5/3}$ for the Goldreich-Sridhar predictions
1003: for the inviscid regime. The fact that the local magnetic perturbations
1004: increase,
1005: albeit slowly, with increasing wavenumber implies that they will eventually
1006: exceed the strength of the background field. Further
1007: research should clarify whether this is a problem.
1008:
1009: %In our model the background field plays no essential
1010: %role, and similar simulations with no imposed large scale field relax
1011: %to a very similar state \citep{SMCM02}.
1012:
1013: If we compare this model to the simulations shown in CLV02b and the
1014: analysis in \citet{CLV03b} we see that the magnetic field Fourier power spectrum
1015: in the simulations is slightly steeper than expected, i.e.
1016: \be
1017: E_B(k)\propto k^{-1.2\hbox{\ to\ }-1.3},
1018: \ee
1019: while the kinetic energy spectrum shows a stronger deviation
1020: \be
1021: E_v(k)\propto k^{\sim -4.5}.
1022: \ee
1023: The fact that the deviation in the exponent is twice as large for
1024: $E_v$ as for $E_B$ is consistent with equation (\ref{s3}). It remains
1025: only to explain why the magnetic power spectrum is not actually flat. The most
1026: likely explanation is that since the power spectrum amplitude at a given $k$
1027: is the Fourier transform of the correlation function, it represents an
1028: integration of the correlation function over all scales $r$ less than
1029: $\sim k^{-1}$.
1030: Since the energy is distributed logarithmically on all scales between
1031: the damping scale and the dissipation scale, this introduces a
1032: factor $\sim\ln (k_{min}/k)$ to $E_B(k)$.
1033: The simulations have a range $\sim 20$ between the scale of viscous damping
1034: and the scale of resistive damping, so this bends the slope
1035: of $E_B(k)$ downward by roughly the required amount.
1036:
1037: The functional form of $\phi_k$ is somewhat more difficult to
1038: test. In CLV03b we constructed filtered maps of $b$ by separating the
1039: ${\bf b}({\bf k})$ into broad bins. Plotting the cumulative
1040: magnetic energy versus volume for each filtered map gives a
1041: measure of the volume filling fraction as a function of scale.
1042: We show the results in figure 2 (Fig.~4 in CLV03b). Comparing
1043: the concentration of magnetic energy as a function of scale
1044: we see that, as expected, the concentration of magnetic power
1045: increases at smaller scales. Our results may have a slightly shallower
1046: dependence on wavenumber than predicted by equation (\ref{4.6a}),
1047: but this depends on the value of the cumulative fractional magnetic
1048: energy we choose for comparison.
1049:
1050: Finally, we note that the eddies in this regime are not described by
1051: the linear solutions to equations (\ref{memovi}) and (\ref{memovn}).
1052: The transfer of magnetic power from large to small scales
1053: makes these solutions largely irrelevant.
1054:
1055: \subsection{The decoupled, damped regime}
1056:
1057: As we have seen, ion-neutral decoupling may happen either before
1058: or after the turbulent
1059: cascade reaches the viscous damping scale. In both cases the cascade
1060: will show similar behavior in the decoupled and damped regime, although the
1061: details of the onset of this stage will differ. We begin by discussing the
1062: case where decoupling occurs below the viscous damping scale, that is, after
1063: the cascade has entered the stage described in the preceding subsection.
1064: We will discuss the case of early decoupling, $\aleph_c<1$, in the middle of
1065: this subsection. We will conclude this subsection by describing the termination
1066: of the damped regime and the reappearance of a strong turbulent cascade involving
1067: only charged particles.
1068:
1069: When $\aleph_c>1$, decoupling occurs as a two step process.
1070: First, the pressure support
1071: from the neutral particles becomes ineffective and the energy cascade
1072: is reduced by a factor of $(P_i/P_{tot})\sim (n_i/n_{tot})$. At somewhat
1073: smaller scales the difference between the ion and neutral velocities
1074: becomes comparable to $v_k$ and the neutrals can be treated as
1075: a static background. We will rely on the same physical arguments used
1076: successfully in the previous subsection to predict
1077: this part of the spectrum. Currently numerical
1078: simulations address only larger scales.
1079:
1080: The first stage of decoupling occurs when the ambipolar diffusion rate
1081: becomes comparable to $\tau_c^{-1}$ and the neutral particles infiltrate
1082: the zones of intense magnetic field perturbations.
1083: This happens at a wavenumber $k_p$ given by
1084: \be
1085: f_nk_p \hat b_p^2\sim {\rho_i\over t_{in}}{1\over k_p\tau_c},
1086: \label{pdecoup}
1087: \ee
1088: where we use the subscript `p' to denote the pressure decoupling
1089: scale. Combining this criterion with equations (\ref{4.1}) and (\ref{4.5})
1090: we see that
1091: \be
1092: {k_p\over k_c} \sim \aleph_c^{1/3}.
1093: \ee
1094: When $f_n\sim 1$ the streaming of neutrals across the perturbed field lines
1095: will be accompanied by significant dissipation, so that the
1096: turbulent energy cascade is
1097: reduced in amplitude by the factor by which the pressure support is
1098: reduced, that is $\sim \rho_i/\rho$. The volume filling factor
1099: must also drop by this same factor in order to maintain the
1100: condition expressed in equation (\ref{s3}).
1101: % We will model this sharp drop in
1102: %$b_k$ and $\phi_k$ at $k_p$ as a discontinuity. This may be
1103: %unrealistic.
1104: %AL
1105: In our simplified model we will model the sharp drop in
1106: $b_k$ and $\phi_k$ at $k_p$ as a discontinuity, although in
1107: reality we expect a smooth transition.
1108: %
1109: It is important to note that there is no associated
1110: discontinuity in the rms electron density fluctuations. The volume
1111: averaged strength of the magnetic pressure fluctuations drops by $\rho_i/\rho$
1112: as the ions go from providing only a fraction ($\rho_i/\rho$) of the
1113: compensating plasma pressure to supplying all of it. The only
1114: observational signal may be a change in higher order moments of the
1115: scintillation statistics.
1116:
1117: The neutrals and ions decouple entirely when the viscous drag coefficient
1118: becomes comparable to the neutral drag term in the ion force equation. This
1119: sets in when
1120: \be
1121: \rho \nu k^2 \sim {\rho_i\over t_{in}},
1122: \ee
1123: or at
1124: \be
1125: k_d\approx k_c (f_n\aleph_c)^{1/2}.
1126: \label{kd}
1127: \ee
1128: For $k>k_d$ the conservation of energy condition and force balance
1129: equations become
1130: \be
1131: b_k^2 \sim b_c^2{\rho_i\over\rho},
1132: \label{c1}
1133: \ee
1134: and
1135: \be
1136: {\rho_i\over t_{in}} {1\over k\tau_c} \sim k_c \hat b_k^2.
1137: \label{c2}
1138: \ee
1139: This implies
1140: \be
1141: \hat b_k\sim (f_n\aleph_c)^{1/2}\left({k_c\over k}\right)^{1/2} b_c,
1142: \label{bsp3}
1143: \ee
1144: \be
1145: \phi_k\sim {k\over k_c} {\rho_i\over\rho}(f_n\aleph_c)^{-1}
1146: \sim{k\over k_c}{t_{in}\over\tau_c},
1147: \label{fsp3}
1148: \ee
1149: \be
1150: v_k\sim v_c \left({k_c\over k}\right)^{1/2}\left({t_{in}\over\tau_c}\right)^{1/2}.
1151: \label{vsp3}
1152: \ee
1153: Equations (\ref{4.5}), (\ref{kd}) and (\ref{bsp3})
1154: imply a maximum local perturbed field strength, at wavenumbers $\sim k_d$, of
1155: \be
1156: b_{max}\sim b_c(f_n\aleph_c)^{1/4}.
1157: \ee
1158: Under most conditions $b_{max}$ will be only moderately larger than $b_c$.
1159: We note that the volume filling fraction {\it increases} with wavenumber
1160: in this regime. This prediction should be tested using a two fluid code.
1161:
1162: How is picture modified when we consider $\aleph_c<1$? The decoupling condition,
1163: $\aleph\sim 1$, evaluated to obtain equation (\ref{decoup1}), is equivalent to
1164: the pressure decoupling criterion given in equation (\ref{pdecoup}) in this case.
1165: We see from equation (\ref{memo7}) that turbulent motions in the ions
1166: are strongly damped below this limit, so that we are immediately in the
1167: damped, pressure decoupled and dynamically decoupled limit.
1168: Consequently, we expect an immediate drop in $\phi_k$ and $b_k$ at the decoupling
1169: scale. On smaller scales we can invoke force balance and energy conservation
1170: to write:
1171: \be
1172: b_k^2 \sim b'^2{\rho_i\over\rho},
1173: \label{d1}
1174: \ee
1175: and
1176: \be
1177: {\rho_i\over t_{in}} {1\over k\tau'} \sim k' \hat b_k^2,
1178: \label{d2}
1179: \ee
1180: where $\tau'$ is the eddy turn over time at the scale $k'$ defined in
1181: equation (\ref{decoup1}).
1182: Equations (\ref{d1}) and (\ref{d2}) imply
1183: \be
1184: \hat b_k\sim b' \left(f_n{k'\over k}\right)^{1/2}\sim b_c\aleph_c^{1/4}
1185: \left(f_n{k_c\over k}\right)^{1/2},
1186: \label{bsp4}
1187: \ee
1188: \be
1189: \phi_k\sim {\rho_i\over\rho_n}\left({k\over k'}\right)
1190: \sim {\rho_i\over\rho_n}\left({k\over k_c}\right)\aleph_c^{-3/2},
1191: \label{bsp4a}
1192: \ee
1193: and
1194: \be
1195: v_k\sim v' \left({k'\over k}\right)^{1/2}\left({\rho_i\over \rho_n}\right)^{1/2},
1196: \label{vsp4a}
1197: \ee
1198: where
1199: $v'$ and $b'$ are the rms velocities and magnetic field perturbation strengths
1200: at the scale $k'$. The differences between equations (\ref{bsp3})-(\ref{vsp3})
1201: and equations (\ref{bsp4})-(\ref{vsp4a}) are due to the different stirring rates
1202: imposed by the marginally damped eddies in the two different cases.
1203:
1204: The persistence of the energy cascade to very small scales implies that
1205: the magnetic field will revert to a strong turbulent
1206: cascade once collisions with the neutral background are slower than the
1207: magnetic dynamical evolution rate. This sets in when
1208: \be
1209: {k \hat b_k\over (4\pi\rho_i)^{1/2}} \sim t_{in}^{-1}.
1210: \label{onset}
1211: \ee
1212: {}From equations (\ref{bsp3}) and (\ref{bsp4}) we find that the onset of small
1213: scale turbulence is at a wavenumber
1214: \be
1215: k_t\sim k_c \left({\rho_n\over\rho_i}\right)\aleph_c\min[1,\aleph_c^{1/2}].
1216: \label{ktdef}
1217: \ee
1218: In both cases $k_t$ coincides with a filling fraction of order unity, so that
1219: the distinction between $\hat b_k$ and $b_k$ disappears.
1220: At this wavenumber small scale instabilities will cause $k_{\|}$ and $v_k$ to
1221: jump sharply, while $b_k$ drops. There will be a sharp mismatch between the
1222: correlation time of the damped turbulence on slightly larger scales and the
1223: eddy turnover time of the strong small scale turbulence, which will start at
1224: $t_{in}$. This suggests that the turbulent cascade will be driven
1225: intermittently, with a buildup of small scale energy followed by
1226: its rapid release. The duty cycle for this process will be approximately
1227: the ratio of $t_{in}$ to the eddy turn over time for the marginally damped
1228: large scale or
1229: \be
1230: \epsilon\sim {\rho_i\over\rho_n} \min[1,\aleph_c^{-1}].
1231: \label{inter}
1232: \ee
1233: During the active phase of the small scale turbulence the cascade rate will
1234: be
1235: \be
1236: b(k_t)^2/t_{in}\sim \rho_n {v_l^4\over V_A l}\max[1,\aleph_c].
1237: \ee
1238: The corresponding turbulent spectrum will be
1239: \be
1240: v_k\sim v_l \left({v_l\over V_A}\right)^{1/3}(kl)^{-1/3}
1241: \left({\rho_n\over\rho_i}\right)^{1/3}
1242: \max[1,\aleph_c^{1/3}],
1243: \label{vsp4}
1244: \ee
1245: and
1246: \be
1247: k_{\|}\sim {1\over l}(kl)^{2/3}\left({v_l\over V_A}\right)^{4/3}
1248: \left({\rho_i\over\rho_n}\right)^{1/6}f_n^{1/3}
1249: \max[1,\aleph_c^{1/3}],
1250: \label{pp}
1251: \ee
1252: with $b_k\sim (4\pi\rho_i)^{1/2}v_k$.
1253:
1254: These bursts of turbulence will terminate at some small
1255: dissipative scale, due to either ohmic resistivity,
1256: ion viscosity along the magnetic field lines or to
1257: plasma effects in a collisionless medium. The resistive
1258: perpendicular wavenumber is set by the condition
1259: $k_{res}\eta\sim v_k$, or
1260: using equation (\ref{vsp4})
1261: \be
1262: k_{res}\sim l^{-1} \left({v_ll\over\eta}\right)^{3/4}
1263: \left({v_l\rho_n\over V_A\rho_i}\right)^{1/4}
1264: \max[1,\aleph_c^{1/4}].
1265: \ee
1266:
1267: For partially ionized gas in the interstellar medium
1268: \citep[see][for a discussion of idealized interstellar phases]{DL98} the
1269: collision rate
1270: is much less than the ion cyclotron frequency and the ion Larmor radius
1271: is greater than $k_{res}^{-1}$. The condition for damping due to
1272: parallel ion viscosity is
1273: \be
1274: k_{\|}^2 c_n l_i\ge k_{\|}V_A^*,
1275: \ee
1276: where, as before, $V_A^*$ is the Alfv\'en velocity for the ions
1277: alone, and $k_{\|}l_i<1$. In the limit where the plasma is
1278: strongly magnetized and $\rho_i\ll\rho_n$ we expect $V_A^*>c_n$, so
1279: that this criterion is never satisfied. That is, damping due to
1280: parallel transport is instead given by free-streaming along the field lines,
1281: which has an associated damping rate which is always less than the
1282: cascade time scale. We have already noted that the perpendicular
1283: ion viscosity
1284: is, at most, only slightly greater than the resistivity. It follows that
1285: the turbulent cascade is truncated at the Larmor radius, when
1286: $kr_L\sim 1$. In dense plasmas, like stars and
1287: accretion disks, collisions dominate, ion viscosity is negligible,
1288: and the cascade ends at the resistive scale.
1289:
1290: \subsection{Turbulence in the ISM}
1291:
1292: Interstellar medium is turbulent with turbulence spreading over a
1293: range of scales from hundreds of parsec \citep[see][]{LP00,SL01}
1294: to astronomical units (see Spangler 1999)
1295: As \citet{ARS95} pointed out, the fact that interstellar scintillation
1296: suggests a power law spectrum consistent with Kolmogorov turbulence
1297: is already a strong indication that the observed scales are
1298: connected by a power cascade. The lack of any feature clearly
1299: attributable to neutral particle damping can be seen as a counter-argument,
1300: but the model sketched above implies that it will be difficult to
1301: detect such a feature. For example, although the average neutral fraction in
1302: the diffuse ionized medium is uncertain, it probably lies somewhere
1303: between a percent and a few times that. From equation (\ref{fcrit})
1304: we see that the critical neutral fraction
1305: for this gas is on the order of several percent, depending on the
1306: local Mach number, so
1307: $f_n\aleph_c$ is at most of order unity. For low Mach numbers
1308: one might expect this to translate
1309: into an exponential suppression by a factor of order unity,
1310: since the eddy turnover rate will drop with the amplitude
1311: of motions on a given scale, allowing for more effective damping.
1312: However, since the electron density fluctuations will trace the magnetic
1313: pressure
1314: fluctuations caused by pseudo-Alfv\'en modes within the turbulent
1315: cascade, the mean square variance of electron density on small
1316: scales, $k_{large}>k_t$, will look like an extrapolation of the same
1317: quantity from large scales, $k_{small}<\min[k',k_c]$,
1318: but reduced by a factor of
1319: \be
1320: \epsilon \left({\rho\over\rho_i}\right){b(k_{large})^2\over b(k_{small})^2}
1321: \left({k_{large}\over k_{small}}\right)^{2/3}
1322: \sim \left({\rho_i\over\rho_n}\right)^{1/3}\min[1,\aleph_c^{-1/3}].
1323: \ee
1324: This will be a very modest reduction. The scintillation power
1325: spectrum will look like a fairly continuous power law from large
1326: to small scales, but with a range of intermediate scales where
1327: it goes flat (from $k_c$ or $k'$ to $k_t$) before dropping
1328: sharply to an extrapolation of its large scale behavior.
1329: The dynamic range of this flat region will be
1330: $\sim (\rho_n/\rho_i)\max[1,\aleph_c]$. Again, in the diffuse
1331: ionized medium this will correspond to a factor of order unity
1332: (or less) in length scales, which will not produce an observational
1333: signature. The warm neutral medium should show a more
1334: pronounced shoulder in the power spectrum, although given
1335: the heterogeneity of the local interstellar
1336: medium, detecting the signal of neutral damping using electron density fluctuations
1337: still represents a challenge. The coldest, and densest, phases of the ISM
1338: have a negligible impact on the scintillation measurements.
1339:
1340: In Table~1 we show critical scales and parameters for some idealized phases
1341: of the ISM, taken from \citet{DL98}. The most obvious point is that the
1342: different regimes discussed in this section cover a very modest range
1343: of scales. Ignoring the parallel scales, which may not leave a clear
1344: observational signal, the entire range of damped scales covers only two
1345: orders of magnitude in the cold and warm neutral phases of the ISM.
1346: For the bulk of the ISM, the turbulent cascade
1347: consists largely of a large scale cascade involving all the particles, and
1348: a smaller scale cascade involving only the ions. (We have not considered the
1349: role of charged grains in this analysis.) Current simulations of strongly
1350: damped MHD turbulence, which apply only to the range between $k_c$ and $k_p$,
1351: already have {\it more} dynamic range than in the ISM itself. The small
1352: wavelength limit on ISM turbulence, set by the Larmor radius, is only
1353: weakly dependent on local conditions.
1354:
1355: In magnetically dominated environments, e.g. in molecular clouds, compressibility
1356: can be important. This will extend the range of scales over which the turbulence
1357: will lie in the viscosity damped regime.
1358:
1359: \section{Reconnection Rates}
1360:
1361: We are now in a position to rederive the speed of reconnection
1362: in turbulent magnetized plasmas, including the effects of a neutral
1363: component. Below we briefly consider highly ionized plasmas, extending
1364: our previous analysis and highlighting steps in our derivation that
1365: will be affected by neutral particles. In \S 3.2 we consider stochastic
1366: reconnection in turbulent plasmas with a small ionized fraction.
1367:
1368: \subsection{Highly ionized plasmas}
1369:
1370: For $f_n<f_{crit}$, as defined in equation (\ref{fcrit}),
1371: the ion-neutral coupling is dynamically insignificant on all scales.
1372: At sufficiently small scales a modest fraction of the energy cascade will
1373: go into heating the neutral gas, but given the approximate nature of
1374: our discussion we will ignore the resulting corrections to
1375: $v_k$. This limit is equivalent to the fully ionized case,
1376: described in LV99. Here we will briefly review the major results
1377: from that paper, since we need to generalize them in order to
1378: calculate reconnection speeds for partially neutral plasmas.
1379:
1380: The basic geometric constraint on reconnection speeds is
1381: \be
1382: V_{rec}={\Delta(L)\over L} V_{eject},
1383: \label{v1}
1384: \ee
1385: where $\Delta(L)$ is the width of the ejection surface for a current
1386: sheet of length $L$. In the absence of viscosity or neutral friction
1387: the ejection speed is $\sim V_A$. The Sweet-Parker rate comes from
1388: taking $\Delta\sim \eta/V_{rec}$. In a collisionless plasma
1389: $\Delta$ can be $\sim r_L$, the ion Larmor radius, when this is larger
1390: than the resistive layer. Both of these
1391: estimates ignore the role that current sheet instabilities may play in
1392: increasing $\Delta(L)$. Here we are principally concerned with another
1393: effect, which is that as the ejected plasma moves out of the reconnection
1394: region, along magnetic field lines, the size of $\Delta(L)$ will
1395: increase as the magnetic fields lines diffuse away from the current sheet.
1396:
1397: In the presence of noise individual magnetic field lines will stay within
1398: the current sheet for some distance, $L_{min}\ll L$, whose value will
1399: depend on the current sheet thickness and the amplitude of the noise.
1400: These individual patches of flux will reconnect at some speed
1401: $V_{rec,local}$ obtained by substituting $L_{min}$ in place of
1402: $L$ in equation (\ref{v1}). In the presence of noise, the current sheet
1403: contains many independent flux elements, all reconnecting simultaneously,
1404: so that we have an upper limit on the reconnection speed given by
1405: $\sim (L/L_{min}) V_{rec,local}$. Since the geometric constraint
1406: on the reconnection speed can be applied on any scale between
1407: $L$ and $L_{min}$ this implies
1408: \be
1409: V_{rec}< {\Delta(L')L\over (L')^2} V_{eject}(L'),\hskip 1cm \hbox{for all\ }
1410: L'\in[L_{min},L],
1411: \label{v2}
1412: \ee
1413: where $\Delta(L')$ is the distance field lines wander perpendicular to the
1414: large scale magnetic field direction within a distance $L'$ along the
1415: field lines. For a turbulent cascade the diffusion length $\Delta$ is
1416: the solution to
1417: \be
1418: {d \Delta^2\over dx} = D_B(\Delta)
1419: \label{v3}
1420: \ee
1421: where $D_B$ defines the stochastic diffusion of field lines separated
1422: by a distance $\Delta$. Assuming that the field line stochasticity is
1423: caused by turbulence in the medium, the functional form of $D_B$ depends on the
1424: nature of the turbulence, and is likely to be a function of scale.
1425: For randomly diffusing field lines this is
1426: \be
1427: D_B(\Delta)\sim \max[\left({b_k\over B_0}\right)^2k_{\|}^{-1}]
1428: \hskip 1cm \hbox{for all\ }k\Delta\ge 1.
1429: \label{diff1}
1430: \ee
1431: where $k_{\|}(k)$ is the parallel wavenumber as a function of the
1432: perpendicular wavenumber. In the presence of strong turbulence this
1433: expression is $k^{-2}k_{\|}(k)\propto k^{-4/3}$.
1434: For long wavelength perturbations, $k\Delta\le1$,
1435: there can be a contribution which is reduced from its value at the scale
1436: $k^{-1}$ by factor $(k\Delta)^2$. Assuming strong turbulence and taking
1437: $\Delta k\sim 1$ we see from equations (\ref{v3}) and (\ref{diff1}) that
1438: we expect $\Delta\propto k_{\|}^{-3/2}\sim L'^{3/2}$. In other words, the
1439: rms separation of field lines should grow as the distance along the direction
1440: of the mean field to the $3/2$ power. In figure 3 we plot rms separations
1441: for a variety of initial separations in a $256^3$ cubed hyperviscosity
1442: and hyperresistivity simulation. The details of the simulation can be
1443: found in CLV02a. We see that the separation grows as distance along
1444: the field lines to the $3/2$ power, as expected, subject
1445: only to a time offset caused by the finite initial separation.
1446: Also as expected, above the dominant eddy scales the curves roll over into the
1447: usual $L^{1/2}$ law. Independent, but similar, simulations are
1448: described in \citet{SMCM02} who found similar results.
1449:
1450: In LV99, and here, we are concerned with the effect of magnetic field
1451: line diffusion on reconnection. However, this same effect removes
1452: the suppression of thermal conduction perpendicular to large scale field
1453: lines in diffuse magnetized plasmas \citep{MN01,CLHKKM03, MCB03, CM03}.
1454: This has had a dramatic effect on our understanding of the thermal
1455: structure and history of hot plasmas in galaxy clusters, which has not
1456: yet reached a final resolution.
1457:
1458: For an ionized plasma with negligible viscosity, $V_{eject}\sim V_A$ at
1459: all scales. Combining equations (\ref{k_p2m}), (\ref{velm}), (\ref{belm}),
1460: (\ref{v2}) and (\ref{diff1})
1461: we find that the most restrictive limit on $V_{rec}$ comes from $L'\sim L$
1462: and gives
1463: \be
1464: V_{rec}< v_T\min\left[\left({L\over l}\right)^{1/2},
1465: \left({l\over L}\right)^{1/2}\right].
1466: \label{eq:lim2a}
1467: \ee
1468: In a real ionized plasma the viscosity is {\it not} necessarily negligible.
1469: When the particle collision rate is low the
1470: perpendicular viscosity
1471: %\footnote{Viscosity is a tensor in the presence
1472: %of magnetic field. According to Cowley (private communication) turbulent
1473: %dynamo is possible when the parallel viscosity is larger than resistivity.
1474: %In our model of viscosity-dominated turbulence parallel viscosity does not
1475: %play a role.}
1476: will be larger than the resistivity
1477: by a factor of several times the ratio of gas pressure to magnetic
1478: field pressure. Under relevant astrophysical circumstances this will
1479: be factor of order $10$, raising the possibility that viscosity may
1480: reduce magnetic field line diffusion at scales slightly larger than
1481: the scale of resistive dissipation
1482: %\footnote{For high beta plasma
1483: %this could induce viscosity-dominated turbulence even in fully ionized
1484: %gas. However, in reality, the turbulence in most cases will be damped at
1485: %the Larmor radius.}.
1486: However, since the strongest
1487: constraint comes from the largest scales, this will not affect
1488: the constraint given in equation (\ref{eq:lim2a}).
1489:
1490: The claim that this is an actual estimate of the reconnection speed,
1491: rather than an upper limit, follows from the absence of any other
1492: important constraints on reconnection. In this case the most obvious
1493: alternative constraint is
1494: that individual flux elements must reconnect many times
1495: after their initial reconnection. If subsequent reconnection events
1496: are slow, then the current sheet will quickly evolve into a tangled
1497: mass of reconnected magnetic field lines. If the scale of the current
1498: sheet is set by the ion Larmor radius, then this will define the
1499: scale for all secondary reconnection events, and the small scale
1500: reconnection speed will be $\sim V_A$. The case where resistivity
1501: defines the width of the current sheet is more complicated. In LV99
1502: we showed that we can consider secondary reconnection events to be
1503: similar to the large scale reconnection, then used a self-similarity
1504: argument to show that secondary reconnection was fast. In either
1505: case we find that equation (\ref{eq:lim2a}) is, as advertised, the
1506: actual reconnection speed for field lines in turbulent background.
1507:
1508: \subsection{Partially ionized plasmas}
1509:
1510: As we have shown in \S 2, a turbulent cascade in a partially ionized plasma
1511: is considerably more complicated than its counterpart in the fully ionized
1512: case. In particular,
1513: the limit on the reconnection speed expressed in equation (\ref{v2}) no
1514: longer increases monotonically with wavenumber, and equation (\ref{eq:lim2a})
1515: may not be a fair estimate of the reconnection speed. We need to consider
1516: the full range of dynamical regimes, and search for minima in the reconnection
1517: speed limit given by equation (\ref{v2}). This requires estimates for
1518: $V_{eject}$ and $D_B(\Delta)$ over the full range of the turbulent cascade.
1519: We have already seen that for $k<k_c\min[1,\aleph_c^{3/2}]$ the most stringent
1520: limit comes from $L'\sim L$, which gives a fairly generous upper limit
1521: on the reconnection speed.
1522:
1523: We begin by noting that when viscous drag is important, the ejection speed
1524: from a volume of thickness $\Delta$ and length $L'$ is
1525: \be
1526: V_{eject}\approx {V_A^2\tau_c\over L'}\max[(k_c\Delta)^2, (f_n\aleph_c)^{-1}].
1527: \label{eject}
1528: \ee
1529: At large scales, $k<k_c$, the plasma acts as a single fluid and we
1530: have $V_{eject}\approx V_A$. At very small scales, when the ions are
1531: completely uncoupled from the neutrals, we get
1532: $V_{eject}\approx V_A(\rho/\rho_i)^{1/2}$.
1533:
1534: For $k>k_t$ a rescaled turbulent cascade
1535: emerges. As long as the current sheet thickness is smaller than $k^{-1}$,
1536: a rescaled version of our previous analysis emerges. In particular,
1537: we expect the most important constraint on the reconnection speed to
1538: arise from the smallest wavenumbers in this regime, $k\sim k_t$. Due
1539: to the intermittent nature of the turbulence we have a factor $\epsilon$
1540: multiplying the usual expression. Using equations
1541: (\ref{ktdef}) and (\ref{inter}) we get
1542: \be
1543: D_B(k>k_t)\sim \epsilon k_t^{-2}k_{\|,t}\left({k_t\over k}\right)^{4/3}\approx
1544: \left({\rho_i\over\rho_n}\right)^{1/2}\left({t_{in}\over\tau_c}\right)^2
1545: {k_{\|,c}\over k_c^2}\left({k_t\over k}\right)^{4/3},
1546: \label{dbd1}
1547: \ee
1548: which is always much less than the diffusion coefficient associated with
1549: the viscous damping scale. However, since it applies to much smaller
1550: scales it still has physical significance.
1551:
1552: The diffusion coefficient for the damped scales is less obvious.
1553: The first complication is that since we are interested in the width
1554: of the ejection zone, the use of an rms value for $\Delta$ is not
1555: obviously correct. The median value would be more appropriate.
1556: Since we are not concerned with constants of order unity here
1557: the distinction is unimportant if the distribution function for
1558: field line separations is a smooth function with a single peak.
1559: Although it is not obvious that this condition is satisfied in
1560: the highly intermittent viscously damped regime, we see from
1561: figure 4 that the median and rms values of field lines separation
1562: seem to track one another as a function of distance along
1563: the field lines. The median is a factor of few lower.
1564:
1565: The second complication is that our model of the viscously damped
1566: regime does not predict the value of $D_B(k)$. Since the curvature
1567: wavenumber of the field lines is constant as a function of scale, we
1568: can estimate the diffusion coefficient as
1569: \be
1570: D_B(k)\sim \left({k_{\|,c}\over k_c}\right)^2l_c \phi_k
1571: \label{dbv}
1572: \ee
1573: where $l_c$ is the distance a field line remains within an intermittent
1574: structure and $\phi_k$ is, as before, the filling factor of such structures.
1575: The value of $l_c$ is uncertain, but goes to $k_{\|,c}^{-1}$ for $k=k_c$,
1576: and must fall much more steeply than $k^{-1/3}$ to be consistent with
1577: the sharply reduced value of $D_B(k)$, relative to the undamped case, seen
1578: in figure 3. Based on a simple geometric picture, we will assume
1579: \be
1580: l_c\approx k_{\|,c}^{-1}{k_c\over k}.
1581: \label{lc}
1582: \ee
1583: This implies that
1584: \be
1585: D_B(k)\sim {k_{\|,c}\over k_c^2}\left({k_c\over k}\right)^2,
1586: \ee
1587: in the damped, viscously coupled regime with $\aleph_c>1$.
1588: However, this is also the level
1589: of diffusion we would expect from the shear imposed at $k=k_c$. Consequently,
1590: the sharp drop in $\phi_k$ at the scale of pressure decoupling does not
1591: affect field line diffusion. In this regime the value of $\Delta$
1592: rises exponentially, with a constant $L'\sim k_{\|,c}^{-1}$. We see in
1593: figure 4 that the simulations are consistent with comparable contributions
1594: to the diffusion coming from large and small scales, and with an
1595: exponential rise in $\Delta$. The sharpest
1596: constraint on $V_{rec}$ will come from the minimum $\Delta$ in this regime.
1597:
1598: Our claim that $l_ck_{\|,c}\ll 1$ is based on the fact that the growth
1599: in field line separations is not consistent with the idea that a small
1600: fraction of field lines diverge sharply from their neighbors for a
1601: distance $\sim k_{\|,c}^{-1}$. We can get some idea of the true situation
1602: by considering the full distribution of field line separations, given in
1603: figure 5. We see that while the separations certainly do not show a
1604: gaussian distribution, they do show that field lines join the growing tail of
1605: large separations at a distance which is typically shorter than $k_{\|,c}^{-1}$.
1606:
1607: Equations (\ref{dbv}) and (\ref{lc}) imply that for $k>k_d$ the
1608: diffusion coefficient will go to a constant value, which persists
1609: until $k=k_t$ and $\phi_k=1$. In this regime, $k_d<k<k_t$, we can
1610: use equation (\ref{ktdef}) and find
1611: \be
1612: D_B(k)\approx {k_{\|,c}\over \min[k_t,k']k_c}\approx
1613: {k_{\|,c}\over k_c^2}{t_{in}\over\tau_c}\max[1,\aleph_c^{-1}],
1614: \label{dbd}
1615: \ee
1616: However,
1617: this value is only relevant when, for some $k>k_d$,
1618: it is greater than the contribution from the
1619: large scale shear given by $\sim k_{\|,c}\Delta^2$ (for $\aleph_c>1$)
1620: or $\sim k_{\|}'\Delta^2$ (for $\aleph_c<1$).
1621: Once this value of $D_B$ applies we get $\Delta\propto L'^{1/2}$ and
1622: the limit on $V_{rec}$ will increase with $k$. The minimum upper
1623: limit on $V_{rec}$ is at a scale
1624: \be
1625: \Delta \approx k_c^{-1} \left({t_{in}\over\tau_c}\right)^{1/2}
1626: \max[1,\aleph_c^{-1/2}],
1627: \label{dmin1}
1628: \ee
1629: with a corresponding $L'$ of
1630: \be
1631: L'\approx k_{\|,c}^{-1}\max[1,\aleph_c^{-1}].
1632: \label{dmin1a}
1633: \ee
1634: We see from equation (\ref{eject}) that the corresponding ejection
1635: velocity is
1636: \be
1637: V_{eject}\approx {V_A\over f_n}\min[1,\aleph_c^{-1}]
1638: \label{eject1}
1639: \ee
1640:
1641: Comparing equations (\ref{dbd1}) and (\ref{dbd}) we that there is a
1642: drop in $D_B(k)$ at $k=k_t$, implying that on some smaller scale
1643: where the larger scale shearing is comparable to the turbulent
1644: diffusion we have another local minimum in the upper limit for $V_{rec}$.
1645: This happens at a field line separation of
1646: \be
1647: \Delta\approx k_t^{-1}\left({t_{in}\over\tau_c}\right)^{3/2}
1648: \left({\rho_i\over\rho}\right)^{3/4}\min[1,\aleph_c^{3/2}]
1649: \approx k_c^{-1}\left({t_{in}\over\tau_c}\right)^{5/2}
1650: \left({\rho_i\over\rho_n}\right)^{3/4}\min[1,\aleph_c],
1651: \label{dmin2}
1652: \ee
1653: with a corresponding $L'$ of
1654: \be
1655: L'\approx k_{\|,c}^{-1} \left({t_{in}\over\tau_c}\right).
1656: \label{dmin2a}
1657: \ee
1658: On such small scales the ejection speed will be $V_A(\rho/\rho_i)^{1/2}$.
1659:
1660: We are now in a position to evaluate the upper limits on the reconnection
1661: speed. We begin with the one set at very small scales. Using equations
1662: (\ref{k_p2m}), (\ref{tau2m}), (\ref{v2}), (\ref{dmin2}),
1663: and (\ref{dmin2a}) we find that
1664: \be
1665: V_{rec}<v_l{L\over l}
1666: \left({\rho_i\over\rho}\right)^{1/4}f_n^{-3/4}
1667: \left({l\over v_lt_{in}}\right)^{1/2} \min[1,\aleph_c^{-1}].
1668: \label{llow}
1669: \ee
1670: This is generally less restrictive than the limit given in equation (\ref{eq:lim2a}).
1671:
1672: The limit on the reconnection speed from intermediate scales can be
1673: found from equations (\ref{4.2.1}), (\ref{4.2.3}), (\ref{als}),
1674: (\ref{v2}), (\ref{dmin1}),
1675: (\ref{dmin1a}), and (\ref{eject1}). In this case we find
1676: \be
1677: V_{rec}<v_T{L\over l}
1678: {\rho_i\over\rho}f_n^{-2}
1679: \left({l\over V_At_{in}}\right)^{1/2} \min[\aleph_c^{-2},\aleph_c^{1/2}].
1680: \label{lim2b}
1681: \ee
1682: This is potentially more restrictive than equation (\ref{eq:lim2a}), and
1683: certainly more restrictive than equation (\ref{llow}), and needs
1684: to be evaluated for specific circumstances. This limit applies whenever
1685: the current sheet is narrower than the value of $\Delta$ given
1686: in equation (\ref{dmin1}). For the specific examples we consider
1687: in the next section this is usually the case.
1688:
1689: \section{Reconnection in Various Phases of ISM}
1690:
1691: All common phases of the ISM are strongly collisionless, so that
1692: ambipolar damping has a reasonable chance of changing the turbulent
1693: power spectrum and consequently the stochastic reconnection speed.
1694: The ISM is extremely heterogeneous, but we can illustrate the
1695: effects of turbulence using the same idealized
1696: phases from \citet{DL98}. We have evaluated the reconnection speed
1697: using equations (\ref{als}) and (\ref{lim2b}) for each phase.
1698: Our results are given in the second to last line of Table~1, and we give a detailed
1699: discussion below. (The last line includes the effects of tearing modes,
1700: which appear to dominate in the densest parts of molecular clouds).
1701: The turbulence which creates
1702: field line stochasticity is assumed to be supplied at some large
1703: scale, specified below, so we are implicitly ignoring the possibility
1704: that as reconnection proceeds it will provide a local source of
1705: %AL
1706: %turbulence.
1707: turbulence\footnote{The increase of stochasticity due to reconnection
1708: may be the source of the finite time instability associated with
1709: solar flares \citep{lv99,VL00}.}
1710: We have also ignored the possibility that reconnection can
1711: change the ionization balance near the current sheet.
1712: %AL
1713: %Finally,
1714: %we have assumed that ions and neutral atoms have roughly the same
1715: %weight, even though this is certainly wrong in the densest
1716: %phases of the ISM, since our calculation is only good to within
1717: %factors of order unity anyhow.
1718:
1719:
1720: We see from an inspection of Table~1 that neutral drag produces
1721: a limit on the reconnection speed which is usually more restrictive than
1722: the limit given at $L'=L$, that is, the limit which applies
1723: to an ionized plasma. The only exception is the warm ionized
1724: medium, where the neutral content is $\sim 1$\% and the turbulent
1725: cascade continues down to $r_L$. Assuming
1726: strong turbulence, with a Mach number of order unity, the reduction
1727: in the reconnection speed from the ionized limit is only a factor
1728: of $\sim10$ for neutral atomic gas. This is moderately sensitive to the scale
1729: of turbulent energy injection, but since we expect that $L\sim l$
1730: this will not be large effect.
1731: On the other hand, strongly subsonic turbulence
1732: implies a much smaller limit on $V_{rec}$. Quiescent regions
1733: in the ISM should have reconnection speeds much smaller than the
1734: local Alfv\'en speed.
1735:
1736: Molecular gas occupies only a small fraction of the ISM's volume,
1737: but plays a special role as the site of star formation. Table~1
1738: indicates that as we consider increasingly dense molecular gas,
1739: we move towards the case of decoupling above the viscous damping
1740: scale. In this limit the reconnection speed becomes more sensitive to
1741: the energy injection scale. We have assumed a single (large)
1742: scale here, but it may be that a much smaller value is appropriate
1743: in small dense regions. In that case the limit on the reconnection
1744: speed would drop significantly. However,
1745: if the density increases to the point where resistive instabilities
1746: produce current sheets substantially broader than the width
1747: given in equation (\ref{dmin1}) then the limit on the reconnection
1748: speed will once again increase. This accounts for the double
1749: entry for molecular clouds. Following the
1750: approach used in other phases we get that
1751: the reconnection speed is lower than
1752: the Alfv\'en speed, even for a Mach number of 1, by more than
1753: an order magnitude.
1754: Unlike the atomic phases, this result scales as the square root of the
1755: turbulent velocity, so that weaker turbulence has a less dramatic effect on the
1756: reconnection speed.
1757: However, this result can be a substantial underestimate. A large scale current
1758: sheet is unstable to tearing modes (\cite{FKR63}), and these provide a localized
1759: source of turbulence which will broaden the current sheet. This sets a minimum
1760: value of the current sheet width. The resulting limit on $V_{rec}$ takes
1761: precedence when it gives a larger $\Delta$, and a larger $V_{rec}$, than
1762: estimates based purely on the turbulent cascade.
1763:
1764: We show the effect of tearing modes in the last line of Table~1.
1765: For $\aleph_c<1$, for example, in the dense cores of molecular clouds, we can
1766: take the parallel correlation scale as
1767: \be
1768: k_{\|}'={1\over V_A}{\rho_i\over\rho t_{in}}.
1769: \ee
1770: The corresponding current sheet width, set by tearing mode instabilities is
1771: \be
1772: \Delta\sim {1\over k_{\|}'}\left({\eta\rho_i\over t_{in}\rho V_A^2}\right)^{3/10},
1773: \ee
1774: for the width of the broadened current sheet. This result is based on the
1775: tearing mode dispersion relation for a current sheet embedded in an
1776: neutral substrate \citep{Z89}, and is equation (\ref{app})
1777: in the appendix.
1778: For motion over scales comparable to $k'^{-1}_{\|}$ we have
1779: $V_{eject}\sim V_A$ so equation (\ref{v2})
1780: implies
1781: \be
1782: V_{rec}\le V_A k_{\|}'L
1783: \left({\eta\rho_i\over t_{in}\rho V_A^2}\right)^{3/10}
1784: \approx V_A\left({\eta\over LV_A}\right)^{3/10}
1785: \left({\rho_i\over\rho}{L\over V_At_{in}}\right)^{13/10}
1786: =V_A\left({V_AL\over\eta_{ambi}}\right)\left({\eta\over\eta_{ambi}}\right)^{3/10},
1787: \ee
1788: where $\eta_{ambi}$ is the ambipolar diffusion coefficient defined by
1789: \be
1790: \eta_{ambi}\equiv {\rho\over\rho_i} V_A^2 t_{in},
1791: \ee
1792: and $\eta<\eta_{ambi}<V_AL$.
1793:
1794: In a dense neutral medium the electron collision
1795: rate is dominated by collisions with neutrals, for which \citet{DRD83}
1796: give
1797: \be
1798: t_{en}^{-1}\approx 2.62\times10^{-9} n_n\left({T_e\over 10 K}\right)^{1/2}.
1799: \ee
1800: Consequently the resistivity for DC phase of the ISM is approximately
1801: $9.3\times10^9$ cm$^2$ sec$^{-1}$. Combining these results we obtain the
1802: last entry of Table~1. The large coefficient implies that the usual large
1803: scale limit, given in equation (\ref{eq:lim2a}), will usually apply here.
1804: In the densest and most neutral phases of the ISM we recover the fast reconnection
1805: speeds typical of an ionized plasma.
1806:
1807: For $\aleph_c>1$ we need to consider tearing modes in a viscous medium, with a parallel
1808: wavenumber $\sim k_{\|,c}$. In this case we find
1809: \be
1810: \Delta\sim k_{\|,c}^{-1}\left({\eta k_{\|,c}\over V_A}\right)^{3/16}
1811: \left({\nu k_{\|,c}\over V_A}\right)^{3/16},
1812: \ee
1813: which is equation (\ref{app2}) in the appendix. Using this expression, and remembering that
1814: $k_c^2\nu=k_{\|,c}V_A$, we obtain
1815: \be
1816: V_{rec}\le V_A (k_{\|,c}L)\left({k_c\over k_{\|,c}}\right)^{7/8}\left({\eta k_{\|,c}\over V_A}\right)^{1/8}.
1817: \ee
1818: The resistivity for our idealized molecular cloud phase is $\sim 1.3\times10^8$. The
1819: resultant value of the reconnection speed in molecular clouds is competitive with the
1820: limit derived from turbulence, but is slightly smaller (meaning that the larger limit
1821: applies). Evidently tearing modes become rapidly more important as we go to
1822: the densest regions of molecular clouds, and are unimportant outside of such regions.
1823:
1824: %For diffuse and dense
1825: % H$_2$ we assume that reconnection and the energy injection
1826: %happens at the scale of 1~pc. As $l_0\approx 10^{13}-10^{12}$~cm
1827: %and $l_{min}\sim l_0 (l/l_0)^{1/2}$ it is possible to show that
1828: %the shearing rate associated with the local outflow of material
1829: %is larger than the recombination rate. Therefore, the local reconnection
1830: %is to happen at corrected Sweet-Parker rate. The criterion given
1831: %by eq.~(\ref{cond1}) is fulfilled and therefore the actual reconnection
1832: %speed is given by (\ref{eq:lim2a}) and is comparable to the Alfv\'en
1833: %speed.
1834:
1835: %Reconnection in mostly ionized media, e.g. warm ionized media is fast
1836: %as it was shown in LV99. For warm diffuse HI the reconnection is
1837: %slowest (of the order of a few cm per second) and this is the
1838: %consequence of the situation when the ambipolar diffusion is suppressed
1839: %while the damping of turbulence is still large.
1840:
1841: Finally, we note that for collisional fluids the magnetic Prandtl number,
1842: $\nu/\eta$, is typically very small, so that the turbulent
1843: cascade extends down to resistive scales. In this case
1844: equation (\ref{eq:lim2a}) is our best estimate of the reconnection
1845: speed, assuming, as before, that $v_l\le V_A$. As an example we can
1846: consider the temperature minimum in the solar photosphere. Adopting
1847: $T\approx 4200$, $n_e\approx 10^{11}$ cm$^{-3}$, and
1848: $n_n\approx 10^{15.2}$ cm$^{-3}$, we find that
1849: $\nu\sim c_n^2t_n\approx 2\times10^5$cm$^2$sec$^{-1}$, and
1850: $\eta\sim 3\times10^9$cm$^2$sec$^{-1}$.
1851:
1852: \section{Conclusions}
1853:
1854: Our results can be summarized as follows:
1855:
1856: \begin{enumerate}
1857: \item
1858: In plasmas characterized by a moderate magnetic Prandtl number
1859: (resistivity less than, or of order, viscosity) magnetic reconnection follows
1860: the model described in LV99, so that reconnection
1861: is fast in the presence of noise. Collisionless plasmas
1862: with small neutral fractions and collisionally dominated
1863: plasmas are typically in this class. In either case
1864: the turbulent cascade proceeds to either the resistive scale or
1865: the ion Larmor radius.
1866: For typical interstellar conditions, it is the ion Larmor radius
1867: which is larger.
1868:
1869: \item
1870: As the neutral content
1871: in plasma increases the turbulent cascade is interrupted at a scale greater than
1872: the neutral mean free path. However, the magnetic field perturbations
1873: are not suppressed below this scale. On the contrary, we find that the
1874: magnetic field power spectrum flattens out, and exhibits more power on
1875: small scales than the Goldreich-Sridhar spectrum. Velocity fluctuations
1876: in this regime are driven by magnetic field perturbations, whose energy is much
1877: larger than the kinetic energy perturbations on the same scale.
1878: The importance of this
1879: new regime of magnetic stochasticity goes far beyond understanding the reconnection
1880: problem.
1881:
1882: \item
1883: On somewhat smaller scales, once the ions and neutrals are
1884: decoupled, the turbulent cascade can reappear. The warm atomic gas
1885: in the ISM shows only a slight interruption in the turbulent
1886: cascade and the small scale scintillations are close to a continuation
1887: of the large scale power. However, this break will be more conspicuous
1888: in colder and denser phases of the ISM. In this limit the small
1889: scale turbulence will be strongly intermittent.
1890:
1891: \item
1892: Magnetic stochasticity on scales smaller than the ion-neutral damping
1893: cut-off of the Goldreich-Sridhar spectrum promotes fast reconnection.
1894: The diffusion of magnetic field lines initially decreases rapidly below this
1895: scale, but the decoupling of ions and neutrals leads to a minimum value
1896: of the diffusion coefficient on small scales. The most stringent limit on
1897: reconnection speeds comes from considering the largest scales where this minimum value
1898: holds. This is in the range of the turbulent cascade where ions and
1899: neutrals are decoupled, but neutral drag is still sufficient to stabilize
1900: magnetic perturbations. In most phases of the ISM this gives a reconnection speed
1901: which is just slightly smaller than the ionized plasma estimate
1902: given in LV99.
1903:
1904: \item
1905: Our study shows that the reconnection speed in the ISM is
1906: always much faster than the predictions given by the Sweet-Parker rate.
1907: Even in very dense and cold regions it is only slightly below the
1908: Alfv\'en speed, and tearing mode instabilities in the current sheet tend
1909: to drive it back up to the Alfv\'en speed.
1910: This process needs to be considered
1911: in models for magnetic flux removal
1912: during star formation, and for studies of the dynamics of magnetized molecular
1913: clouds.
1914: \end{enumerate}
1915:
1916: Finally, we note that our conclusions are moderately sensitive to the
1917: nature of field line diffusivity in the various damped regimes
1918: ($k_t<k<\min[k',k_c]$). We have argued that the shearing due to
1919: motions at $k\sim k_c$ dominate until we are well into the decoupled
1920: damped regime. Our estimate for a constant field line diffusivity
1921: on somewhat smaller scales has not been tested against simulations.
1922: If this turns out to be too high (or low) then our reconnection
1923: speed estimate will have to be lowered (or raised).
1924:
1925:
1926:
1927: \acknowledgements
1928:
1929: AL and JC acknowledge NSF grant AST-0125544 and ETV acknowledges
1930: the support of the NSF grant AST-0098615. ETV would also like to
1931: thank the hospitality of the organizers of the Festival de Theorie
1932: in Aix-le-Provence, 2001, where some of this work was done.
1933:
1934: \appendix
1935: \section{Tearing Modes and Current Sheet Width}
1936:
1937: The current sheet implicit in the Sweet-Parker picture of magnetic
1938: reconnection is well known to be unstable to tearing modes
1939: \citep{FKR63}, which will create a turbulent zone between the
1940: volumes of unreconnected magnetic flux.
1941: In LV99 we suggested that the broadening of the current
1942: sheet implied by this effect should lead to an enhanced reconnection
1943: rate for laminar field lines, on the order of
1944: \be
1945: V_{rec}\approx V_A\left({\eta\over V_A L}\right)^{3/10}.
1946: \ee
1947: Broadly similar results were obtained by \citet{S88}, except that we
1948: have calculated the typical size of the magnetic field fluctuations
1949: using tearing mode theory rather than leaving it as a free parameter.
1950:
1951: \citet{Z89} derived
1952: the dispersion relation for tearing modes in
1953: a current sheet of thickness $\Delta$ embedded in a neutral substrate.
1954: It is
1955: \be
1956: {1\over (k_{\|}\Delta^2)^4}={\gamma^5\rho_i\over k_{\|}^2V_A^2\rho}
1957: {\Delta^2\over\eta^3}\left(1+{t_{in}^{-1}\over\gamma+(\rho_i/\rho)t_{in}^{-1}}\right),
1958: \label{a1}
1959: \ee
1960: where $\gamma$ is the tearing mode growth rate and
1961: $k_{\|}\Delta<<1$. This favors modes with minimal $k_{\|}$, i.e. transverse
1962: wavelengths comparable to the current sheet length. The resulting
1963: turbulence will widen the current sheet until the tearing modes are marginally
1964: stabilized by the shear due to the ejection of plasma, i.e.
1965: $\gamma\approx V_A/L\approx k_{\|}V_A$.
1966:
1967: When the ions and neutrals are decoupled the ejection velocity from the current
1968: sheet is
1969: \be
1970: V_{eject}\approx k_{\|}V_A^2{\rho\over\rho_i}t_{in}.
1971: \ee
1972: The dispersion relation for $\gamma\ge (\rho_i/\rho)t_{in}^{-1}$ is
1973: \be
1974: \gamma^4={V_A^2\eta^3 t_{in}\over \Delta^{10} k_{\|}^2}.
1975: \ee
1976: So for the saturated modes
1977: \be
1978: \Delta k_{\|}\approx \left({\eta\rho_i\over V_A^2\rho t_{in}}\right)^{3/10}.
1979: \label{app}
1980: \ee
1981:
1982: This determines the width of the outflow zone if we can assume that
1983: the tearing modes produce stochastic mixing of the field lines, that is, that
1984: neighboring field lines are well-mixed within this volume. In two dimensions
1985: this is obviously not the case. In three dimensions it is probably valid.
1986: We are also assuming that the existence of turbulence in the current sheet
1987: does not lead to widespread mixing of field lines outside the current sheet.
1988: Although some mixing seems inevitable, the fact that the modes are driven
1989: only as long as they are narrow enough to fit within the current sheet,
1990: and grow only about as fast as plasma, and magnetic structures, are ejected
1991: from the current sheet, suggests that any induced field line mixing will
1992: only spread the the outflow zone by an additional factor of order unity.
1993:
1994: When the ions and neutrals are {\it not} decoupled, but viscous drag
1995: plays an important role we can rewrite equation (\ref{a1}) as
1996: \be
1997: \left({\Delta'\over I}\right)^4={\gamma^5\rho_i\over k_{\|}^2V_A^2\rho}
1998: {\Delta^2\over\eta^3}\left(1+{(\nu/\Delta^2)\over\gamma}\right),
1999: \ee
2000: where we have assumed a single fluid ($\gamma> (\rho_i/\rho)t_{in}^{-1}$)
2001: and replaced drag by a neutral substrate with viscous damping.
2002: The ejection velocity in this case is obtained by balancing viscous drag
2003: with magnetic forces,
2004: \be
2005: k_{\|}V_A^2\sim {\nu\over \Delta^2}V_{eject}.
2006: \ee
2007: We conclude that in this case
2008: \be
2009: \Delta\sim k_{\|,c}^{-1}\left({\eta k_{\|,c}\over V_A}\right)^{3/16}
2010: \left({\nu k_{\|,c}\over V_A}\right)^{3/16}.
2011: \label{app2}
2012: \ee
2013:
2014: %\section{Reconnection Induced by Fast Modes}
2015: %
2016: %According to Cho \& Lazarian (2002, 2003) the fast modes exhibit the
2017: %isotropic spectrum similar to that of acoustic turbulence. Unlike
2018: %the Goldreich-Sridhar
2019: %turbulence case when the amplitude of waves $v_k/\omega_A$
2020: %is proportional to $k^{-1}$, in acoustic turbulence it is proportional
2021: %to $k^{-1} v_k/V_A\ll k^{-1}$.
2022: %If this is
2023: %true, then the treatment of reconnection provided in the Appendix~D
2024: %of LV99 should be applicable. According
2025: %to it the local reconnection rates are
2026: %\be
2027: %V_{rec, local}\sim V_A \left(\frac{v_l}{V_A}\right)^{2/3}\left(\frac{\eta}{V_A l}\right)^{1/6},
2028: %\ee
2029: %which are higher than those induced by Alfvenic turbulence.
2030: %
2031: \begin{thebibliography}{}
2032: \bibitem[Armstrong, Rickett \& Spangler(1997)]{ARS95}
2033: Armstrong, J.W., Rickett, B.J. \& Spangler, S.R. 1995,\apj, 443, 209
2034: \bibitem[Bhattacharjee \& Hameiri(1986)]{BH86}Bhattacharjee, A. \&
2035: Hameiri, E.\ 1986, \prl, 57, 206
2036: \bibitem[Bhattacharjee, Ma, \& Wang(2001)]
2037: {BMW01} Bhattacharjee, A., Ma, Z.W. \& Wang, X. 2001, Phys. Plasmas, 8(5), 1829
2038: \bibitem[Bhattacharjee, Ma, \& Wang(2003)]
2039: {BMW03} Bhattacharjee, A., Ma, Z.W. \& Wang, X. 2003,
2040: in {\it Turbulence and Magnetic Fields in Astrophysics},
2041: eds. T. Passot \& E.
2042: Falgarone (Springer Lecture Notes in Physics 614; 2003),
2043: 351
2044: \bibitem[Biskamp(1996)]{B96}Biskamp, D.\ 1996, Astrophys. \& Sp. Sci., 242,165
2045: \bibitem[Biskamp(2000)]{B00}Biskamp, D.\ 2000, Magnetic Reconnection
2046: in Plasmas (Cambridge: Cambridge University Press)
2047: %\bibitem[Biskamp, Drake \& Chen(1994)]{BDC94}
2048: %Biskamp, D., Drake, J.F. \& Chen, J. 1994, J. Geophys. Res.,
2049: %95, 18833
2050: \bibitem[Biskamp, Schwarz \& Drake(1997)]{BSD97}
2051: Biskamp, D., Schwarz, E. \& Drake, J.F. 1997, Phys. Plasmas,
2052: 4, 1002
2053: \bibitem[Brandenburg(2001)]{B01}Brandenburg, A.\ 2001, \apj, 550, 824
2054: \bibitem[Cattaneo \& Hughes(1996)]{CH96}Cattaneo, F., \& Hughes, D.W.\ 1996,
2055: \pre, 54, 4532
2056: \bibitem[Chandran \& Maron(2003)]{CM03}Chandran, B.D.G. \& Maron, J.L.
2057: \ 2003, astro-ph/0303214, submitted to \prl
2058: \bibitem[Cho \& Lazarian(2002)]{CL02}
2059: Cho, J. \& Lazarian A. 2002, Phys. Rev. Lett., 88, 245001
2060: \bibitem[Cho \& Lazarian(2003)]{CL03}
2061: Cho, J. \& Lazarian A. 2003, MNRAS, 345, 325
2062: \bibitem[Cho \& Vishniac(2000)]{CV00}
2063: Cho, J., \& Vishniac, E. T. 2000, \apj, 538, 217
2064: \bibitem[Cho, Lazarian \& Vishniac(2002a)]{CLV02a}
2065: Cho, J., Lazarian, A. \& Vishniac, E. T. 2002a, \apj, 564, 291
2066: \bibitem[Cho, Lazarian \& Vishniac(2002b)]{CLV02b}
2067: Cho, J., Lazarian, A. \& Vishniac, E. T. 2002b, \apj, 566, 49L
2068: \bibitem[Cho, Lazarian \& Vishniac\ 2003a]{CLV03a}
2069: Cho, J., Lazarian, A. \& Vishniac, E. T. 2003a,
2070: in {\it Turbulence and Magnetic Fields in Astrophysics},
2071: eds. T. Passot \& E.
2072: Falgarone (Springer Lecture Notes in Physics 614; 2003),
2073: 56
2074: \bibitem[Cho, Lazarian \& Vishniac(2003b)]{CLV03b} Cho, J., Lazarian, A., \&
2075: Vishniac, E.T. \ 2003b, \apj, 595, 812
2076: \bibitem[Cho et al.(2003)]{CLHKKM03}Cho, J., Lazarian, A., Honein, A.,
2077: Knaepen, B., Kassinos, S. \& Moin, P.\ 2003, \apj, 589L, 77
2078: \bibitem[Dere(1996)]{D96} Dere, K.P. 1996, \apj, 472, 864
2079: \bibitem[Draine, Roberge \& Dalgarno(1983)]{DRD83}Draine, B.T., Roberge,
2080: W.G., \& Dalgarno, A. 1983, \apj, 264, 485
2081: \bibitem[Draine \& Lazarian(1998)]{DL98}
2082: Draine, B.T., \& Lazarian, A. 1998, \apj, 494, L19
2083: \bibitem[Furth, Killeen, \& Rosenbluth(1963)]{FKR63} Furth, H.P.,
2084: Killeen, J., \& Rosenbluth, M.N.\ 1963, Phys. Fluids, 6, 459
2085: \bibitem[Goldreich \& Sridhar(1995)]{gs95} Goldreich, P. \& Sridhar, S.\ 1995,
2086: \apj, 438, 763 (GS95)
2087: \bibitem[Gruzinov \& Diamond(1994)]{GD94}Gruzinov, A.V. \& Diamond, P.H.\ 1994,
2088: \prl, 72, 1651
2089: \bibitem[Gruzinov \& Diamond(1996)]{GD96}Gruzinov, A.V. \& Diamond, P.H.\ 1996,
2090: Phys. Plasmas, 3, 1853
2091: %\bibitem[Gruzinov(1998)]{G98} Gruzinov, A.V.\ 1998, \apj, 501, 787
2092: \bibitem[Hameiri \& Bhattacharjee(1987)]{HB87}Bhattacharjee, A. \&
2093: Hameiri, E.\ 1987, Phys. Fluids, 30, 1744
2094: \bibitem[Heitsch \& Zweibel(2003a)]{HZ03a}Heitsch, F. \& Zweibel, E.G.\ 2003a,
2095: \apj, 583, 229
2096: \bibitem[Heitsch \& Zweibel(2003b)]{HZ03b}Heitsch, F. \& Zweibel, E.G.\ 2003b,
2097: \apj, 590, 291
2098: \bibitem[Hughes et al.(1996)]{HCK96}
2099: Hughes, D.W., Cattaneo, F. \& Kim, E.J.\ 1996, Phys. Lett. A, 223, 167
2100: \bibitem[Innes, Inhester, Axford \& Wilhelm(1997)]{IIAW97}
2101: Innes, D.E., Inhester, B., Axford, W.I., \& Wilhelm, K. 1997, Nature, 386, 811
2102: \bibitem[Jacobson \& Moses(1984)]{JM84}Jacobson, A.R. \& Moses, R.W. 1984,
2103: Phys. Rev. A, 29(6), 3335
2104: \bibitem[Ji, Yamada, Hsu \& Kulsrud(1998)]{JYHK98}Ji, H., Yamada, M., Hsu,
2105: S. \& Kulsrud, R.\ 1998, \prl, 80, 3256
2106: \bibitem[Kim \& Diamond(2001)]{KD01} Kim, E.-J., \& Diamond, P.H.\ 2001,
2107: \apj, 556, 1052
2108: \bibitem[Krause \& Radler(1980)]{KR80} Krause, F., \& Radler, K.H. 1980,
2109: Mean-Field Magnetohydrodynamics and Dynamo Theory (Oxford: Pergamon Press)
2110: \bibitem[Kulsrud \& Pearce(1969)]{KP69} Kulsrud, R., \& Pearce, W.P. 1969,
2111: \apj, 156,445
2112: \bibitem[Lazarian \& Pogosyan(2000)]{LP00}
2113: Lazarian, A. \& Pogosyan, D.\ 2000, \apj, 537, 720
2114: %\bibitem[Lazarian \& Vishniac(1996)]{LV96}
2115: %Lazarian A., \& Vishniac, E.T.\ 1996, in Polarimetry of the
2116: %Interstellar Medium, eds. W.G.~Roberge and D.C.B.~Whittet, ASP 97, 537
2117: \bibitem[Lazarian \& Vishniac(1999)]{lv99}
2118: Lazarian, A. \& Vishniac, E.T. 1999, \apj, 517, 700 (LV99)
2119: \bibitem[Lazarian \& Vishniac(2000)]{lv00} Lazarian, A. \& Vishniac, E.T.
2120: \ 2000, Rev.Mex.~de~Astron.~y~Astrof., 9, 55
2121: \bibitem[Lithwick \& Goldreich(2001)]{LG01}Lithwick, Y. \& Goldreich, P.
2122: \ 2001, \apj, 562, 279
2123: \bibitem[Maron \& Goldreich(2001)]{MG01}
2124: Maron, J. \& Goldreich, P. 2001, \apj, 554, 1175
2125: \bibitem[Maron, Chandran \& Blackman(2003)]{MCB03}Maron, J.L., Chandran,
2126: B.D.G. \& Blackman, E.G.\ 2003, astro-ph/0303217
2127: \bibitem[Matthaeus \& Lamkin(1985)]{ML85} Matthaeus, W.H. \& Lamkin, S.L.
2128: \ 1985, Phys. Fluids, 28, 303
2129: \bibitem[McIvor(1977)]{M77}McIvor, I. 1977, \mnras, 178, 85
2130: \bibitem[Minter \& Spangler(1997)]{MS97}Minter, A.H., \& Spangler, S.R.\ 1997,
2131: \apj, 485, 182
2132: \bibitem[Moffatt(1978)]{M78} Moffatt, H.K. 1978, Magnetic Field Generation in E
2133: lectrically
2134: Conducting Fluids (Cambridge: Cambridge University Press)
2135: \bibitem[Naidu, McKenzie \& Axford(1992)]{NMA92}
2136: Naidu, K., McKenzie, J.F., \& Axford, W.I. 1992, Ann.
2137: Geophys., 10, 827
2138: \bibitem[Narayan \& Medvedev (2001)]{MN01}
2139: Narayan, R. \& Medvedev, M.V.\ 2001, \apj, 562L, 129
2140: %\bibitem[Narayan \& Yi(1995)]{NY95}
2141: %Narayan, R. \& Yi, I. 1995, \apj, 452, 710
2142: \bibitem[Parker(1957)]{P57}Parker, E.N.\ 1957, J. Geophys. Res., 62, 509
2143: \bibitem[Parker(1979)]{P79}\rule{1.2cm}{0.2mm}\ 1979, Cosmical
2144: Magnetic Fields (Oxford: Clarendon Press)
2145: \bibitem[Parker(1992)]{P92}Parker, E.N.\ 1992, \apj, 401, 137
2146: \bibitem[Petschek(1964)]{P64}Petschek, H.E.\ 1964,
2147: {\it The Physics of Solar Flares}, AAS-NASA
2148: Symposium, NASA SP-50 (ed. W.H. Hess), Greenbelt, Maryland, p.~425
2149: \bibitem[Priest \& Forbes(2000)]{PF00}Priest, E. \& Forbes, T.\ 2000,
2150: Magnetic Reconnection: MHD Theory and Applications (Cambridge: Cambridge
2151: University Press)
2152: \bibitem[Schekochihin, Maron, Cowley \& McWilliams(2002)]{SMCM02}
2153: Schekochihin, A., Maron, J., Cowley, S. \& McWilliams, J.\ 2002, \apj,
2154: 576, 806
2155: \bibitem[Schekochihin, Cowley, Maron \& McWilliams(2003)]{SCMM03}
2156: Schekochihin, A., Cowley, S., Maron, J. \& McWilliams, J.
2157: \ 2003, astro-ph/0308336
2158: \bibitem[Shay \& Drake(1998)]{SD98}
2159: Shay, M.A., Drake, J.F. 1998, Geophys. Res. Let., 25(20), 3759
2160: \bibitem[Shay, Drake, Denton \& Biskamp(1998)]{SDD98}
2161: Shay, M.A., Drake, J.F., Denton, R.E., \& Biskamp, D.
2162: 1998, J. Geophys. Res., 103, 9165
2163: \bibitem[Spangler(1991)]{S91} Spangler,S.R. \ 1991,\apj, 376, 540
2164: \bibitem[Spangler(1999)]{Sp99} Spangler,S.R. \ 1999,\apj, 522, 879
2165: \bibitem[Speiser(1970)]{S70}Speiser, T.W. 1970, Planet. Space Sci., 18,
2166: 613
2167: \bibitem[Spitzer(1978)]{S78} Spitzer, L.\ 1978, Physical Processes in the
2168: Interstellar Medium (New York: John Wiley \& Sons)
2169: \bibitem[Stanimirovic \& Lazarian(2001)]{SL01}Stanimirovic, S \& Lazarian, A.\ 2001, \apj, 441, 53
2170: \bibitem[Strauss(1985)]{S85} Strauss, H.R.\ 1985, Phys. Fluids, 28, 2786
2171: \bibitem[Strauss(1988)]{S88} Strauss, H.R.\ 1988, \apj, 326, 412
2172: \bibitem[Sweet(1958)]{S58} Sweet, P.A.\ 1958, in IAU Symp. 6, Electromagnetic
2173: Phenomena in
2174: Cosmical Plasma, ed. B. Lehnert (New York: Cambridge Univ. Press), 123
2175: \bibitem[Trintchouk, Yamada, Ji, Kulsrud \& Carter(2003)]{TYJKC03}
2176: Trintchouk, F., Yamada, M., Ji, H., Kulsrud, R.M. \& Carter, T.A.\ 2003,
2177: Physics of Plasmas, 10(1), 319
2178: \bibitem[Vainshtein \& Cattaneo(1992)]{VC92}Vainshtein, S.I. and Cattaneo, F.
2179: \ 1992, \apj, 393, 165
2180: %\bibitem[Vishniac(1995a)]{V95a} Vishniac, E.T.\ 1995a, \apj, 446, 724
2181: %\bibitem[Vishniac(1995b)]{V95b}\rule{1.2cm}{0.2mm}\ 1995b, \apj, 451, 816
2182: \bibitem[Vishniac \& Lazarian(1999)]{VL99}
2183: Vishniac, E.T. \& Lazarian, A.\ 1999, \apj, 511, 193
2184: \bibitem[Vishniac \& Lazarian(2000)]{VL00}
2185: Vishniac, E.T., \& Lazarian, A. 2000, in
2186: {\it Plasma Turbulence and Energetic Particles},
2187: ed. by M. Ostrowski, R. Schlickeiser
2188: (Cracow, 2000) p.182
2189: \bibitem[Vishniac \& Cho(2001)]{VC01}
2190: Vishniac, E.T. \& Cho, J.\ 2001, \apj, 550, 752
2191: \bibitem[Yan \& Lazarian(2002)]{YL02}
2192: Yan, H. \& Lazarian, A. 2002, Phys. Rev. Lett, 89, 281102
2193: \bibitem[Zweibel(1989)]{Z89}Zweibel, E.G.\ 1989, \apj, 340, 550
2194: %\bibitem[Zweibel(1998)]{Z98} Zweibel, E.G. 1998, Physics of Plasmas, v. 5, 247
2195: \bibitem[Zweibel \& Brandenburg(1997)]{ZB97}
2196: Zweibel, E.G., \& Brandenburg, A., 1997, \apj, 478, 563
2197: \end{thebibliography}
2198: \clearpage
2199:
2200: \begin{deluxetable}{lccccc}
2201: \tablecolumns{6}
2202: \rotate
2203: \setlength{\tabcolsep}{0.02in}
2204: \tabletypesize{\scriptsize}
2205: \tablecaption{Scales and Reconnection speeds for idealized phases of the ISM}
2206: \tablehead{
2207: \colhead{ISM:} & \colhead{WIM} &
2208: \colhead{WNM} & \colhead{CNM} &
2209: \colhead{MC} & \colhead{DC}}
2210: \startdata
2211: $n$&$0.1$&$0.4$&$30$&$300$&$10^4$\\[0.4cm]
2212: $x$&$0.99$&$0.1$&$10^{-3}$&$10^{-4}$&$10^{-6}$\\[0.4cm]
2213: $T$&$8000$&$6000$&$100$&$20$&$10$\\[0.4cm]
2214: $F_M$&$0.14$&$0.43$&$0.43$&\nodata&\nodata\\[0.4cm]
2215: $\aleph_c$&$9.1l_{30}^{1/2}{\cal M}^{-1}\beta^{1/4}$&$16l_{30}^{1/2}{\cal M}^{-1}\beta^{1/4}$
2216: &$7.8l_{30}^{1/2}{\cal M}^{-1}\beta^{1/4}$
2217: &$4.8l_{30}^{1/2}{\cal M}^{-1}\beta^{1/4}$&$0.37l_{30}^{1/2}{\cal M}^{-1}\beta^{1/4}$\\[0.4cm]
2218: $k_c^{-1}$&\nodata&$2.4\times10^{16}l_{30}^{1/4}{\cal M}^{-1/2}\beta^{-1/8}$
2219: &$3.4\times10^{15}l_{30}^{1/4}{\cal M}^{-1/2}\beta^{-1/8}$
2220: &$8.8\times10^{13}l_{30}^{1/4}{\cal M}^{-1/2}\beta^{-1/8}$
2221: &\nodata\\[0.4cm]
2222: \tableline
2223: $k_{\|,c}^{-1}$&\nodata&$3.8\times10^{17}l_{30}^{1/2}{\cal M}^{-1}\beta^{-1/4}$
2224: &$1.0\times10^{17}l_{30}^{1/2}{\cal M}^{-1}\beta^{-1/4}$
2225: &$9.0\times10^{15}l_{30}^{1/2}{\cal M}^{-1}\beta^{-1/4}$
2226: &\nodata
2227: \\[0.4cm]
2228: $k'^{-1}$&\nodata&\nodata&\nodata&\nodata&$2.6\times10^{13}l_{30}^{-1/2}{\cal M}\beta^{-1/2}$\\
2229: [0.4cm]
2230: $k_{\|}^{\prime -1}$&\nodata&\nodata&\nodata&\nodata
2231: &$4.0\times10^{15}\beta^{-1/2}$\\
2232: [0.4cm]
2233: $k_p^{-1}$&\nodata&$9.6\times10^{15}l_{30}^{1/12}{\cal M}^{-1/6}\beta^{-5/24}$
2234: &$1.7\times10^{15}l_{30}^{1/12}{\cal M}^{-1/6}\beta^{-5/24}$
2235: &$5.2\times10^{13}l_{30}^{1/12}{\cal M}^{-1/6}\beta^{-5/24}$
2236: &\nodata
2237: \\[0.4cm]
2238: $k_d^{-1}$&\nodata&$6.4\times10^{15}\beta^{-1/4}$&$1.2\times10^{15}\beta^{-1/4}$
2239: &$4.0\times10^{13}\beta^{-1/4}$&\nodata\\[0.4cm]
2240: \tableline
2241: $k_t^{-1}$&\nodata&$1.7\times10^{14}{\cal M}^{1/2}l_{30}^{-1/4}\beta^{-3/4}$&
2242: $1.3\times10^{13}{\cal M}^{1/2}l_{30}^{-1/4}\beta^{-3/4}$&
2243: $1.8\times10^9{\cal M}^{1/2}l_{30}^{-1/4}\beta^{-3/4}$&
2244: $2.6\times10^7{\cal M}l_{30}^{-1/2}\beta^{-1/2}$
2245: \\[0.4cm]
2246: $k_{\|,t}^{-1}$&\nodata&
2247: $8.5\times10^{15}\beta^{-1/2}$
2248: &$1.3\times10^{14}\beta^{-1/2}$&
2249: $1.9\times10^{13}\beta^{-1/2}$
2250: &$4.0\times10^{15}\beta^{-1/2}$\\[0.4cm]
2251: $r_L$&$7.3\times10^7\beta^{1/2}$&$1.2\times10^8\beta^{1/2}$&
2252: $1.3\times10^{8}\beta^{1/2}$&$1.3\times10^8\beta^{1/2}$&
2253: $2.3\times10^8\beta^{1/2}$\\[0.4cm]
2254: $V_{rec}$&$v_T{L_{30}^{1/2}\over l_{30}^{1/2}}\min[1,l_{30}/L_{30}]$&
2255: $0.092v_T{\cal M}^2\beta^{-1/2}{L_{30}\over l_{30}^{3/2}}$&
2256: $0.077v_T{\cal M}^2\beta^{-1/2}{L_{30}\over l_{30}^{3/2}}$&
2257: $0.095v_T{\cal M}^2\beta^{-1/2}{L_{30}\over l_{30}^{3/2}}$&
2258: $0.094v_T{\cal M}^{-1/2}\beta^{3/4}{L_{30}\over l_{30}^{1/4}}$\\[0.4cm]
2259: $V_{rec}(TM)$&\nodata&\nodata&\nodata&$0.05v_T{\cal M}^{-17/8}l^{9/16}_{30}L_{30}\beta^{-61/64}$&
2260: $21v_T{\cal M}^{-1}\beta^{-0.3}L_{30}$
2261: \enddata
2262: \tablecomments{Here ${\cal M}$ is the turbulence Mach number,
2263: $\equiv v_T/c_n$, $\beta\equiv c_n^2/V_A^2$, $F_M$ is the (approximate)
2264: mass fraction contained in each phase,
2265: and a subscript `$30$' indicates that distances are given in
2266: units of $30$ parsecs. The second row gives the
2267: ionization fraction for each phase. The abbreviations in the first row
2268: denote `Warm Ionized Medium', `Warm Neutral Medium', `Cold Neutral Medium',
2269: `Molecular Cloud', and `Dense Core' (in a molecular
2270: cloud). Length scales are defined in the text and are given here in
2271: centimeters.}
2272: \end{deluxetable}
2273: \clearpage
2274:
2275: %\begin{deluxetable}{lccccc}
2276: %\tablewidth{40pc}
2277: %\tablecaption{Reconnection speeds for idealized phases of the ISM}
2278: %\tablehead{
2279: %\colhead{ISM phase} & \colhead{$n$} &
2280: %\colhead{$x$} & \colhead{$T$} &
2281: %\colhead{$\aleph_c$} & \colhead{$V_{rec}$}}
2282: %\startdata
2283: %WIM&$0.1$&$0.99$&$8000$&$9.1l_{30}^{1/2} {\cal M}^{-1}
2284: %\left({c_n\over V_A}\right)^{1/2}$&
2285: %$1.1\times10^3 v_T{\cal M}^2\left({V_A\over c_n}\right)^{1/2}
2286: %{L_{30}\over l_{30}^{3/2}}$\\
2287: %\hline
2288: %WNM&$0.4$&$0.1$&$6000$&$1.6\times10l_{30}^{1/2} {\cal M}^{-1}
2289: %\left({c_n\over V_A}\right)^{1/2}$&
2290: %$9.2\times10^{-2} v_T{\cal M}^2\left({V_A\over c_n}\right)^{1/2}
2291: %{L_{30}\over l_{30}^{3/2}}$\\
2292: %\hline
2293: %CNM&$30$&$10^{-3}$&$100$&$7.8l_{30}^{1/2} {\cal M}^{-1}
2294: %\left({c_n\over V_A}\right)^{1/2}$&
2295: %$7.7\times 10^{-2}v_T {\cal M}^2\left({V_A\over c_n}\right)^{1/2}
2296: %{L_{30}\over l_{30}^{3/2}}$\\
2297: %\hline
2298: %MC&$300$&$10^{-4}$&$20$&$4.8l_{30}^{1/2} {\cal M}^{-1}
2299: %\left({c_n\over V_A}\right)^{1/2}$&
2300: %$9.5\times10^{-2} v_T{\cal M}^2\left({V_A\over c_n}\right)^{1/2}
2301: %{L_{30}\over l_{30}^{3/2}}$\\
2302: %\hline
2303: %DC&$10^4$&$10^{-6}$&$10$&$0.37l_{30}^{1/2} {\cal M}^{-1}
2304: %\left({c_n\over V_A}\right)^{1/2}$&
2305: %$9.4\times10^{-2} v_T{\cal M}^{-1/2}\left({c_n\over V_A}\right)^{3/4}
2306: %{L_{30}\over l_{30}^{1/4}}$\\
2307: %TM&"&"&"&"&$21V_A \left({c_n\over V_A}\right)^{1.6}L_{30}$\\
2308: %\hline
2309: %\enddata
2310: %\tablecomments{Here ${\cal M}$ is the turbulence Mach number,
2311: %$\equiv v_T/c_n$,
2312: %and a subscript `$30$' indicates that distances are given in
2313: %units of $30$ parsecs. The second column gives the
2314: %ionization fraction for each phase. The abbreviations in the first column
2315: %denote `Warm Ionized Medium', `Warm Neutral Medium', `Cold Neutral Medium',
2316: %`Molecular Cloud', `Dense Core', and `Tearing Modes' (in a dense molecular
2317: %cloud core).}
2318: %\end{deluxetable}
2319: %\clearpage
2320:
2321:
2322: \begin{figure}
2323: \plotone{f1.eps}
2324: \caption{Upper plot: Sweet-Parker scheme of reconnection. Middle plot:
2325: illustration of stochastic reconnection that accounts for field line
2326: noise. Lower plot: a close-up of the contact region.
2327: Thick arrows depict outflows of plasma. From \cite{lv00}.}
2328: \end{figure}
2329:
2330: \clearpage
2331:
2332: \begin{figure}
2333: \plotone{f2.eps}
2334: \caption{Fractional volume (X-axis) vs.
2335: fractional magnetic energy in the volume (Y-axis)
2336: for viscosity-damped MHD turbulence:
2337: smaller scales show a higher concentration of magnetic energy.
2338: }
2339: \end{figure}
2340:
2341: \clearpage
2342:
2343: \begin{figure}
2344: \epsscale{.60}
2345: \plotone{f3.eps}
2346: \caption{Root mean square separation of field lines in a simulation of
2347: inviscid MHD turbulence, as a function of
2348: distance parallel to the mean magnetic field, for a range of initial
2349: separations. Each curve represents 1600 line pairs.
2350: The simulation has been filtered to remove pseudo-Alfv\'en
2351: modes, which introduce noise into the diffusion calculation.}
2352: \end{figure}
2353:
2354: \clearpage
2355:
2356: \begin{figure}
2357: \epsscale{.40}
2358: \plotone{f4a.eps}
2359: \plotone{f4b.eps}
2360: \plotone{f4c.eps}
2361: \caption{Root mean square and median separation of field lines in simulation of
2362: viscously damped MHD turbulence, as a function of distance parallel to
2363: the mean magnetic field. In (a) we see the total result. In (b) we have
2364: removed long wavelength modes. In (c) we retain only the long wavelength
2365: modes. In all three cases the median tracks the mean, at a slightly
2366: smaller amplitude, and is consistent with exponential growth. The simulation
2367: used a $384^3$ grid. The data are based on 1600 line pairs.}
2368: \end{figure}
2369:
2370: \clearpage
2371:
2372: \begin{figure}
2373: \plotone{f5.eps}
2374: \caption{A histogram of field line separations for viscously damped
2375: turbulence. The distribution shown here is the same one used for the
2376: curve in figure 4a for an initial separation of 4 grid points.
2377: The tail of large separations grows smoothly and
2378: continuously, and involves a large fraction of all field lines despite
2379: the intermittency of the magnetic field perturbations.}
2380: \end{figure}
2381: \end{document}
2382:
2383:
2384:
2385:
2386: