1: \documentclass[12pt,preprint]{aastex}
2: \def\bm#1{\mbox{\boldmath $#1$}} %--- vector (bold face)
3: \usepackage{pslatex}
4: \usepackage{graphicx}
5:
6: \begin{document}
7:
8: \title{Strong Anisotropic MHD Turbulence with Cross Helicity}
9: \author{Benjamin D. G. Chandran}
10: \email{benjamin.chandran@unh.edu}
11: \affil{Space Science Center and
12: Department of Physics, University of New Hampshire}
13:
14: \begin{abstract}
15: This paper proposes a new phenomenology for strong incompressible MHD
16: turbulence with nonzero cross helicity. This phenomenology is then
17: developed into a quantitative Fokker-Planck model that describes the
18: time evolution of the anisotropic power spectra of the fluctuations
19: propagating parallel and anti-parallel to the background magnetic
20: field~$\bm{B}_0$. It is found that in steady state the power spectra
21: of the magnetic field and total energy are steeper than
22: $k_\perp^{-5/3}$ and become increasingly steep as $C/{\cal E}$
23: increases, where $C=\displaystyle \int d^3\! x \; \bm{v}\cdot\bm{B}$
24: is the cross helicity, ${\cal E}$ is the fluctuation energy, and
25: $k_\perp$ is the wavevector component perpendicular to~$\bm{B}_0$.
26: Increasing $C$ with fixed ${\cal E}$ increases the time required for
27: energy to cascade to smaller scales, reduces the cascade power,
28: and increases the anisotropy of the small-scale
29: fluctuations. The implications of these results for the solar wind and
30: solar corona are discussed in some detail.
31: \end{abstract}
32: \keywords{turbulence --- magnetic fields --- magnetohydrodynamics
33: --- solar wind --- solar corona --- solar flares}
34:
35: \maketitle
36:
37: \section{Introduction}
38:
39: Much of our current understanding of incompressible
40: magnetohydrodynamic (MHD) turbulence has its roots in the pioneering
41: work of Iroshnikov~(1963) and Kraichnan~(1965). These studies
42: emphasized the important fact that Alfv\'en waves travelling in the
43: same direction along a background magnetic field do not interact with
44: one another and explained how one can think of the cascade of energy
45: to small scales as resulting from collisions between oppositely
46: directed Alfv\'en wave packets. They also argued that in the absence
47: of a mean magnetic field, the magnetic field of the energy-containing
48: eddies at scale~$L$ affects fluctuations on scales~$\ll L$ much in the
49: same way as would a truly uniform mean magnetic field.
50:
51: Another foundation of our current understanding is the finding that
52: MHD turbulence is inherently anisotropic. Montgomery \& Turner~(1981)
53: and Shebalin, Matthaeus, \& Montgomery~(1983) showed that a strong
54: uniform mean magnetic field~$\bm{B}_0$ inhibits the cascade of energy
55: to small scales measured in the direction parallel to~$\bm{B}_0$. This
56: early work was substantially elaborated upon by Higdon~(1984),
57: Goldreich \& Sridhar~(1995, 1997), Montgomery \& Matthaeus~(1995), Ng
58: \& Bhattacharjee~(1996, 1997), Galtier et al~(2000), Cho \&
59: Lazarian~(2003), Oughton et al~(2006), and many others. For example,
60: Cho \& Vishniac (2000) used numerical simulations to show that when
61: the fluctuating magnetic field~$\delta B$ is $\ga B_0$ the small-scale
62: turbulent eddies become elongated along the local magnetic field
63: direction. Goldreich \& Sridhar (1995) introduced the important and
64: influential idea of ``critical balance,'' which holds that at each
65: scale the linear wave period for the bulk of the fluctuation energy is
66: comparable to the time for the fluctuation energy to cascade to
67: smaller scales. Goldreich \& Sridhar (1995, 1997), Maron \&
68: Goldreich~(2001), and Lithwick \& Goldreich~(2001) clarified a number
69: of important physical processes in anisotropic MHD turbulence and used
70: the concept of critical balance to determine the ratio of the
71: dimensions of turbulent eddies in the directions parallel and
72: perpendicular to the local magnetic field.
73:
74: Over the last several years, research on MHD turbulence has been
75: proceeding along several different lines. For example, one group of
76: studies has attempted to determine the power spectrum, intermittency,
77: and anisotropy of strong incompressible MHD turbulence using direct
78: numerical simulations. (See, e.g., Cho \& Vishniac 2000, M\"uller \&
79: Biskamp 2000, Maron \& Goldreich 2001, Cho et~al~2002, Haugen
80: et~al~2004, Muller \& Grappin~2005, Mininni \& Pouquet 2007, Perez \&
81: Boldyrev 2008). Another series of papers has explored the properties
82: of anisotropic turbulence in weakly collisional magnetized plasmas
83: using gyrokinetics, a low-frequency expansion of the Vlasov equation
84: that averages over the gyromotion of the particles. (Howes et
85: al~2006, 2007a, 2007b; Schekochihin et al 2007). These authors
86: investigated the transition between the Alfv\'en-wave cascade and a
87: kinetic-Alfv\'en-wave cascade at length scales of order the proton
88: gyroradius~$\rho_i$, as well as the physics of energy dissipation and
89: entropy production in the low-collisionality regime. Turbulence at
90: scales~$\lesssim \rho_i$ has also been examined both numerically and
91: analytically within the framework of fluid models, in particular Hall
92: MHD and electron MHD. (Biskamp, Schwarz, \& Drake 1996, Biskamp
93: et~al~1999, Matthaeus et al~2003; Galtier \& Bhattacharjee 2003, 2005;
94: Cho \& Lazarian 2004; Brodin et al~2006, Shukla et al~2006). Another
95: group of studies has investigated the power spectrum, intermittency,
96: and decay time of compressible MHD turbulence. (Oughton et al 1995,
97: Stone et~al~1998, Lithwick \& Goldreich~2001, Boldyrev et al~2002,
98: Padoan et al~2004, Elmegreen \& Scalo 2004). Additional work by
99: Kuznetsov (2001), Cho \& Lazarian (2002, 2003), Chandran (2005), and
100: Luo \& Melrose (2006) has begun to address the way in which Alfv\'en
101: waves, fast magnetosonic waves, and slow magnetosonic waves interact
102: in compressible weak MHD turbulence. Another recent development is the
103: finding that strong incompressible MHD turbulence leads to alternating
104: patches of alignment and anti-alignment between the velocity and
105: magnetic-field fluctuations. (Boldyrev 2005, 2006; Beresnyak \&
106: Lazarian 2006; Mason, Cattaneo, \& Boldyrev 2006; Matthaeus
107: et~al~2007) These studies examined how the degree of local alignment
108: (and anti-alignment) depends upon length scale, as well as the effects of
109: alignment upon the energy cascade time and the power spectrum of the
110: turbulence.
111:
112: The topic addressed in this paper is the role of cross helicity
113: in incompressible MHD turbulence. The cross helicity is defined as
114: \begin{equation}
115: C = \int d^3\!x \; \bm{v}\cdot {B},
116: \label{eq:defC}
117: \end{equation}
118: where $\bm{v}$ is the velocity and~$\bm{B}$ is the magnetic field. The
119: cross helicity is conserved in the absence of dissipation and can be
120: thought of as a measure of the linkages between lines of vorticity and
121: magnetic field lines, both of which are frozen to the fluid flow in
122: the absence of dissipation (Moffatt 1978). In the presence of a
123: background magnetic field, $\bm{B}_0 = B_0 \hat{z}$, the cross
124: helicity is also a measure of the difference between the energy of
125: fluctuations travelling in the $-z$ and $+ z$ directions. Dobrowolny,
126: Mangeney, \& Veltri~(1980) showed that MHD turbulence with cross
127: helicity decays to a maximally aligned state, with $\delta \bm{v} =
128: \pm \delta\bm{ B}/\sqrt{4\pi\rho}$, where $\delta \bm{v}$ and $\delta
129: \bm{B}$ are the fluctuating velocity and magnetic field and $\rho$ is
130: the mass density. Different decay rates for the energy and cross
131: helicity were also demonstrated by Matthaeus \& Montgomery~(1980). In
132: another early study, Grappin, Pouquet, \& L\'eorat~(1983) used a
133: statistical closure, the eddy-damped quasi-normal Markovian (EDQNM)
134: approximation, to study strong 3D incompressible MHD turbulence with
135: cross helicity, assuming isotropic power spectra. They found that
136: when~$C\neq 0$, the total energy spectrum is steeper than the
137: isotropic Iroshnikov-Kraichnan $k^{-3/2}$~spectrum. Pouquet
138: et~al~(1988) found a similar result in direct numerical simulations of
139: 2D incompressible MHD turbulence. More recently, Lithwick, Goldreich,
140: \& Sridhar~(2007) and Beresnyak \& Lazarian (2007) addressed the role
141: of cross helicity in strong MHD turbulence taking into account the
142: effects of anisotropy.
143:
144: This paper presents a new phenomenology for strong, anisotropic,
145: incompressible MHD turbulence with nonzero cross helicity, and is
146: organized as follows. Section~\ref{sec:wpc} presents some relevant
147: theoretical background. Section~\ref{sec:theory} introduces the new
148: phenomenology as well as two nonlinear advection-diffusion equations
149: that model the time evolution of the power spectra. Analytic and
150: numerical solutions to this equation in the weak-turbulence and
151: strong-turbulence regimes are presented in Sections~\ref{sec:weak}
152: and~\ref{sec:strong}. Section~\ref{sec:strong} also presents a simple
153: phenomenological derivation of the power spectra and anisotropy in
154: strong MHD turbulence. Section~\ref{sec:trans} presents a numerical
155: solution to the advection-diffusion equation that shows the smooth
156: transition between the weak and strong turbulence
157: regimes. Section~\ref{sec:unequal} addresses the case in which the
158: parallel correlation lengths of waves propagating in opposite
159: directions along the background magnetic field are unequal at the
160: outer scale. In Section~\ref{sec:solarwind}, the proposed
161: phenomenology is applied to turbulence in the solar wind and solar
162: corona, and in Section~\ref{sec:comp} the results of this work are
163: compared to the recent studies of Lithwick, Goldreich, \&
164: Sridhar~(2007) and Beresnyak \& Lazarian~(2007).
165:
166:
167: \section{Energy Cascade from Wave-Packet Collisions}
168: \label{sec:wpc}
169:
170: The equations of ideal incompressible MHD can be written
171: \begin{equation}
172: \frac{\partial \bm{w}^\pm}{\partial t} +\left(
173: \bm{w}^\mp \mp v_{\rm A}\hat{z} \right) \cdot \nabla \bm{w}^\pm
174: = -\nabla \Pi
175: \label{eq:mhd}
176: \end{equation}
177: where $\bm{w}^\pm = \bm{v} \pm (\delta\bm{B}/\sqrt{4\pi \rho})$ are
178: the Elsasser variables, $\bm{v}$ is the fluid velocity, $\delta
179: \bm{B}$ is the magnetic field fluctuation, $\rho$ is the mass density,
180: which is taken to be uniform and constant, $v_A = B_0/\sqrt{4\pi
181: \rho}$ is the Alfv\'en speed, $\bm{B}_0 = B_0 \hat{z}$ is the mean
182: magnetic field, and $\Pi = (p + B^2/8\pi)/\rho$, which is determined
183: by the incompressibility condition, $\nabla \cdot \bm{w}^\pm = 0$.
184: Throughout this paper it is assumed that $\delta B \lesssim B_0$ and
185: $w^\pm \lesssim v_A$.
186:
187: In the limit of small-amplitude fluctuations ($w^\pm \ll v_A$), the
188: nonlinear term $\bm{w}^\mp \cdot \nabla \bm{w}^\pm$ in
189: equation~(\ref{eq:mhd}) can be neglected to a first approximation, and
190: the curl of equation~(\ref{eq:mhd}) becomes
191: \begin{equation}
192: \left(\frac{\partial }{\partial t} \mp v_A \frac{\partial}{\partial
193: z}\right) \nabla \times \bm{w}^\pm = 0,
194: \end{equation}
195: which is solved by setting $\nabla \times \bm{w}^\pm$ equal to an
196: arbitrary function of $z\pm v_A t$. Thus, $\bm{w}^\pm$ represents
197: fluctuations with $\bm{v} = \pm \bm{b}$ that propagate in $\mp z$
198: direction at speed~$v_A$ in the absence of nonlinear interactions. In
199: the absence of an average velocity, the cross helicity defined in
200: equation~(\ref{eq:defC}) can be rewritten as
201: \begin{equation}
202: C = \frac{\sqrt{\pi\rho}}{2} \int d^3\! x \, \left[(w^+)^2 - (w^-)^2\right].
203: \label{eq:defC2}
204: \end{equation}
205: The cross helicity is thus proportional to the difference in energy
206: between fluctuations propagating in the $-z$ and $+z$ directions.
207:
208: Equation~(\ref{eq:mhd}) shows that the nonlinear term is nonzero only
209: at those locations where both $w^+$ and $w^-$ are nonzero. Nonlinear
210: interactions can thus be thought of as collisions between oppositely
211: directed wave packets (Kraichnan 1965). When both $w^+$ and $w^-$ are
212: nonzero, equation~(\ref{eq:mhd}) indicates that the $\bm{w}^\pm$
213: fluctuations are advected not at the uniform velocity $\mp v_A
214: \hat{z}$, but rather at the non-uniform velocity $\mp v_A\hat{z} +
215: \bm{w}^\mp$. Maron \& Goldreich (2001) elaborated upon this idea by
216: showing that to lowest order in fluctuation amplitude, if one neglects
217: the pressure term, then $w^+$ wave packets are advected along the
218: hypothetical magnetic field lines corresponding to the sum of
219: $\bm{B}_0$ and the part of $\delta \bm{B}$ arising from the $w^-$
220: fluctuations. This result can be used to construct a geometrical
221: picture for how wave-packet collisions cause energy to cascade to
222: smaller scales, as depicted in Figure~\ref{fig:f1}. In this figure,
223: two oppositely directed wave packets of dimension $\sim \lambda_\perp$
224: in the plane perpendicular to~$\bm{B}_0$ and
225: length~$\lambda_\parallel$ along~$\bm{B}_0$ pass through one another
226: and get sheared. Collisions between wavepackets of
227: similar~$\lambda_\perp$ are usually the dominant mechanism for
228: transferring energy from large scales to small scales. The duration
229: of the collision illustrated in the figure is approximately~$\Delta t
230: \sim \lambda_\parallel/v_A$. The fluctuating velocity and
231: magnetic field are taken to be in the plane perpendicular
232: to~$\bm{B}_0$, as is the case for linear
233: shear Alfv\'en waves. The magnitude of the nonlinear term in
234: equation~(\ref{eq:mhd}) is then $\sim w^+_{\lambda_\perp}
235: w^-_{\lambda_\perp}/\lambda_\perp$, where $w^\pm_{\lambda_\perp}$ is
236: the rms amplitude of the $w^\pm$ wave packet. The fractional change in
237: the $\bm{v}$ and $\bm{b}$ fields of the $w^\mp$ wave packet induced by
238: the collision is then roughly
239: \begin{equation}
240: \left(\frac{w^+_{\lambda_\perp} w^-_{\lambda_\perp}}{\lambda_\perp}\right)\times
241: \left(\frac{\Delta t}{
242: w^\mp_{\lambda_\perp}}\right) = \frac{w_{\lambda_\perp}^\pm \lambda_\parallel}{v_A \lambda_\perp}.
243: \label{eq:defchi1}
244: \end{equation}
245:
246: \begin{figure}[h]
247: \includegraphics[width=3in]{f1.eps}
248: \caption{\footnotesize When two wave packets collide, each wave packet
249: follows the field lines of the other wave packet and gets sheared.
250: \label{fig:f1}}
251: \end{figure}
252:
253: If this fractional change is $\ll 1$ for both $w^+$ and $w^-$
254: fluctuations then neither wave packet is altered significantly by a
255: single collision, and the turbulence is weak. Wave packets travel a
256: distance~$\gg \lambda_\parallel$ before being significantly distorted,
257: and the fluctuations can thus be viewed as linear waves that are only
258: weakly perturbed by nonlinear interactions with other waves. In the
259: wave-packet collision depicted in Figure~\ref{fig:f1}, the right-hand
260: side of the $ w^-$ wave packet is altered by the collision in almost
261: the same way as the left-hand side, since both sides encounter
262: essentially the same~$w^+$ wave packet, since the~$w^+$ packet is
263: changed only slightly during the collision. Changes to the profile of
264: a wave packet along the magnetic field are thus weaker than changes in
265: the profile of a wave packet in the plane perpendicular to~$\bm{B}_0$
266: (Shebalin et al 1983, Ng \& Bhattacharjee 1997, Goldreich \& Sridhar
267: 1997, Bhattacharjee \& Ng~2001, Perez \& Boldyrev~2008). As a result,
268: in the weak-turbulence limit, the cascade of energy to
269: small~$\lambda_\parallel$ is much less efficient than the cascade of
270: energy to small~$\lambda_\perp$ (Galtier et al 2000).
271:
272: On the other hand, if the fractional change in
273: equation~(\ref{eq:defchi1}) is of order unity then a $w^\mp$ wave
274: packet is distorted substantially during a single collision, and the
275: turbulence is said to be ``strong.'' In the case that the fractional
276: change in equation~(\ref{eq:defchi1}) is $\sim 1$ for one fluctuation
277: type, (e.g., $w^-$) but $\ll 1$ for the other ($w^+$), the turbulence
278: is still referred to as strong. It should be noted that strong
279: turbulence can arise when $w^\pm_{\lambda_\perp} \ll v_A$, provided
280: that $\lambda_\perp \ll \lambda_\parallel$. In strong turbulence
281: energy cascades to smaller~$\lambda_\parallel$ to a greater extent
282: than in weak turbulence, but the primary direction of energy flow
283: in~$k$-space is still to larger~$k_\perp$, as discussed in the next
284: section.
285:
286: \section{Anisotropic MHD Turbulence with Cross Helicity}
287: \label{sec:theory}
288:
289: In order to develop an analytical model, it is convenient to work
290: in terms of the Fourier transforms of the fluctuating $w^\pm$ fields, given by
291: \begin{equation}
292: \tilde{\bm{w}}^\pm(\bm{k}) = \frac{1}{(2\pi)^3}\int d^3\!x \; \bm{w}(\bm{x})
293: e^{-i\bm{k}\cdot \bm{x}}.
294: \end{equation}
295: The three-dimensional power spectrum~$A^\pm(\bm{k})$ is defined by the equation
296: \begin{equation}
297: \langle \tilde{\bm{w}}^\pm (\bm{k}) \cdot \tilde{\bm{w}}^\pm(\bm{k}_1)\rangle
298: = A^\pm(\bm{k}) \delta(\bm{k}+ \bm{k}_1),
299: \end{equation}
300: where $\langle \dots \rangle$ denotes an ensemble average.
301: Cylindrical symmetry about~$\bm{B}_0$ is assumed, so that
302: $A^\pm(\bm{k}) = A^\pm (k_\perp,k_\parallel)$, where
303: $k_\perp$ and $k_\parallel$ are the components of $\bm{k}$
304: perpendicular and parallel to~$\bm{B}_0$.
305: The mean-square velocity associated with $w^\pm$ fluctuations
306: is then
307: \begin{equation}
308: (\delta v^\pm)^2 = \frac{1}{4}\int d^3 \!k \; A^\pm(k_\perp,k_\parallel).
309: \label{eq:defdeltav}
310: \end{equation}
311:
312: It is assumed
313: that at each value of~$k_\perp$ there is a parallel wave number
314: $\overline{k}_\parallel ^{\,\pm}(k_\perp)$ such that
315: (1) the bulk of the $w^\pm$ fluctuation energy is
316: at $|k_\parallel| < \overline{k}_\parallel ^{\,\pm}(k_\perp)$
317: and (2) $A^\pm(k_\perp,k_\parallel)$ depends only weakly
318: on~$k_\parallel$ for $|k_\parallel| < \overline{k}_\parallel ^{\,\pm}(k_\perp)$.
319: A $w^\pm$
320: wavepacket at perpendicular scale $k_\perp^{-1}$ then has a correlation
321: length in the direction of the mean field of $\sim
322: \left(\overline{k}_\parallel^{\,\pm} \right)^{-1}$. The rms amplitude
323: of the fluctuating Elsasser fields at a perpendicular scale
324: $k_\perp^{-1}$, denoted $w_{k_\perp}^\pm$, is given by
325: \begin{equation}
326: (w^\pm_{k_\perp})^2 \sim A^\pm(k_\perp,0)k_\perp^2 \overline{k}_\parallel^{\,\pm}.
327: \label{eq:defdeltakpar}
328: \end{equation}
329: As described in the section~\ref{sec:wpc}, when a $w^\mp$ wave packet
330: at scale $k_\perp^{-1}$ collides with a $w^\pm$ wave packet at scale
331: $k_\perp^{-1}$, the fractional change in the $w^\mp$ packet resulting
332: from the collision is approximately
333: \begin{equation}
334: \chi_{k_\perp}^\pm = \frac{k_\perp w_{k_\perp}^\pm}{{\overline k}_\parallel ^{\,\pm} v_A}.
335: \label{eq:defchi2}
336: \end{equation}
337: The wave number $k_c^\pm$ is defined to be the value of $\overline{k}_\parallel^{\,\pm}$
338: for which $\chi_{k_\perp}^\pm = 1$. Thus,
339: \begin{equation}
340: k_c^\pm = \frac{k_\perp^4 A^\pm(k_\perp, 0)}{v_A^2}.
341: \label{eq:defkc}
342: \end{equation}
343:
344: \subsection{The energy cascade time}
345:
346: When $\overline{k}_\parallel^{\,-} \gg k_c^-$, the value of
347: $\chi_{k_\perp} ^-$ is~$\ll 1$ and a $w^+$ is only weakly affected by
348: a single collision with a $w^-$ wave packet. Each such collision
349: requires a time $(\overline{k}_\parallel^{\,-} v_A)^{-1}$. The
350: effects of successive collisions add incoherently, and thus
351: $(\chi_{k_\perp}^-)^{-2}$ collisions are required for the $w^+$ wave
352: packet to be strongly distorted, and for its energy to pass to smaller
353: scales. The cascade time $\tau_{k_\perp}^+$ for a $w^+$ wave packet at
354: perpendicular scale~$k_\perp^{-1}$ is
355: thus roughly
356: \begin{equation}
357: \tau_{k_\perp}^+ \sim (\overline{k}_\parallel^{\,-} v_A)^{-1} (\chi_{k_\perp}^-)^{-2}
358: \sim \frac{1}{k_c^- v_A} \mbox{ \hspace{0.3cm} (weak turbulence).}
359: \label{eq:tau1}
360: \end{equation}
361: Similarly, if $\chi_{k_\perp}^+ \ll 1$, then $\tau_{k_\perp}^- \sim (k_c^+ v_A)^{-1}$.
362:
363: When $\overline{k}_\parallel^{\,-} \sim k_c^-$, the value of $\chi_{k_\perp}
364: ^-$ is~$\sim 1$, a $w^+$ is strongly distorted during a single
365: wave packet collision, and the turbulence is strong. Each such
366: collision takes a time $(\overline{k}_\parallel^{\,-}
367: v_A)^{-1}$. Since $\overline{ k}_\parallel^{\,-} \sim k_c^-$,
368: \begin{equation}
369: \tau_{k_\perp}^+ \sim (\overline{k}_\parallel^{\,-}
370: v_A)^{-1} \sim \frac{1}{k_c^- v_A} \mbox{ \hspace{0.3cm} (strong turbulence).}
371: \label{eq:tau2}
372: \end{equation}
373: Similarly, if $\chi_{k_\perp}^+ \sim 1$, then $\tau_{k_\perp}^- \sim
374: (k_c^+ v_A)^{-1}$.
375:
376:
377: The case $\overline{k}_\parallel^{\,\pm} \ll k_c^\pm$ (i.e.,
378: $\chi_{k_\perp}^\pm \gg 1$) is explicitly excluded from the
379: discussion. Initial conditions could in principle be set up in which
380: $\overline{k}_\parallel^{\,\pm} \ll k_c^\pm$. However, the cascade
381: mechanisms described in section~\ref{sec:parcasc} will not produce the
382: condition $\overline{k}_\parallel^{\,\pm} \ll k_c^\pm$ if it is not
383: initially present. It should be emphasized that in both weak
384: turbulence and strong turbulence, the cascade time is given by the
385: same formula, $\tau_{k_\perp}^{\pm} \sim (k_c^\mp v_A)^{-1}$, which
386: involves the $A^\mp$ spectrum evaluated only at $k_\parallel=0$.
387:
388: \subsection{The Cascade of Energy to Larger~$k_\parallel$}
389: \label{sec:parcasc}
390:
391: The two basic mechanisms for transferring fluctuation energy to
392: larger~$k_\parallel$ were identified by Lithwick, Goldreich, \&
393: Sridhar~(2007). The
394: first of these can be called ``propagation with distortion.'' Suppose
395: a $w^+$ wave packet of perpendicular scale $k_\perp^{-1}$ and
396: arbitrarily large initial parallel correlation length begins colliding
397: at $t=0$ with a stream of $w^-$ wave packets of similar perpendicular
398: scale. At time $t=\tau_{k_\perp}^+$, the leading edge of the $w^+$
399: wave packet has been distorted substantially by the stream of $w^-$
400: wave packets, but the trailing portion of the $w^+$ wave packet at
401: distances $\ga 2 v_A \tau_{k_\perp}^+$ behind the leading edge has not
402: yet encountered the stream of $w^-$ wave packets. If the parallel
403: correlation length of the~$w^+$ wave packet is initially $ >2 v_A
404: \tau_{k_\perp}^+$, then during a time~$\tau_{k_\perp}^+$ the $w^+$
405: wave packet acquires a spatial variation in the direction of the
406: background magnetic field of length scale~$\sim 2 v_A \tau_{k}^+ \sim
407: 2 (k_c^-)^{-1}$. This process is modeled as diffusion of $w^\pm$
408: fluctuation energy in the $k_\parallel$ direction with diffusion
409: coefficient $D_\parallel^\pm \sim (\Delta k_\parallel)^2/\Delta t$,
410: where $\Delta k_\parallel = k_c^\mp$ and $\Delta t = \tau_k^\pm$.
411: ``Propagation with distortion'' then leads to a value of
412: $D_\parallel^\pm$ of $\sim (k_c^\mp)^3 v_A$.
413:
414: The second mechanism identified by Lithwick, Goldreich, \&
415: Sridhar~(2007) can be called ``uncorrelated cascade.'' Consider a
416: $w^+$ wave packet of perpendicular scale~$k_\perp^{-1}$ and
417: arbitrarily large parallel correlation length, and consider two points
418: within the wave packet, $P_1$ and $P_2$, that move with the wave
419: packet at velocity~$-v_A \bm{\hat{z}}$ and are separated by a distance
420: along~$\bm{B}_0$ of $ 2v_A \tau_{k_\perp}^- \sim 2(k_c^+)^{-1}$. The
421: $w^-$ wave packets at perpendicular scale~$k_\perp^{-1}$ encountered
422: by the portions of the $w^+$ wave packet at $P_1$ and $P_2$ are then
423: uncorrelated, because $w^-$ wave packets are substantially distorted
424: while propagating between $P_1$ and $P_2$. Thus, the way in which the
425: $w^+$ wave packet cascades at location~$P_1$ is not correlated with
426: the way in which the~$w^+$ wave packet cascades at location~$P_2$. If
427: the parallel correlation length of the~$w^+$ wave packet is
428: initially~$>2v_A \tau_{k_\perp}^-$, then wave-packet collisions
429: introduce a spatial variation along~$\bm{B}_0$ into the $w^+$ wave
430: packet of length scale~$\sim 2v_A \tau_{k_\perp}^- \sim 2(k_c^+)^{-1}$
431: during a time~$\tau_{k_\perp}^+$. Again, we model this as diffusion
432: of $w^\pm$ fluctuation energy in the $k_\parallel$ direction with
433: $D_\parallel^\pm \sim (\Delta k_\parallel)^2/\Delta t$ and $\Delta t =
434: \tau_{k_\perp}^\pm$, but now $\Delta k_\parallel = k_c^\pm$.
435: ``Uncorrelated cascade'' thus leads to a $k_\parallel$-diffusion
436: coefficient of~$\sim (k_c^\pm)^2 k_c^\mp v_A$.
437:
438:
439: Accounting for both mechanisms, one can write
440: \begin{equation}
441: D^\pm_\parallel \sim (k_{c,\rm max})^2 k_c^\mp v_A,
442: \label{eq:Dpar}
443: \end{equation}
444: where $k_{c,\rm max}(k_\perp)$ is the larger of $k_c^+(k_\perp)$ and
445: $k_c^-(k_\perp)$. If $k_c^+ > k_c^-$, then $w^+$ energy
446: diffuses in~$k_\parallel$ primarily through the ``uncorrelated
447: cascade'' mechanism, while $w^-$ energy
448: diffuses in~$k_\parallel$ primarily through
449: the ``propagation with distortion cascade'' mechanism.
450:
451: \subsection{Advection-Diffusion Model for the Power Spectra}
452:
453: The phenomenology described in the preceding sections is encapsulated
454: by the following nonlinear advection-diffusion equation,
455: \begin{equation}
456: \frac{\partial A^\pm_k}{\partial t} = - \frac{1}{k_\perp}
457: \frac{\partial }{\partial k_\perp} \left(\frac{c_1 k_\perp^2 A^\pm_k h^\pm_k}{\tau^\pm _{{\rm eff,} k_\perp}}\right) + c_2 (k_{c,\rm max})^2 k_c^\mp v_A \frac{\partial^2 A^\pm_k}{\partial k_\parallel^2} + S^\pm_k - \gamma^\pm_k A^\pm_k,
458: \label{eq:FPpm}
459: \end{equation}
460: where $A^\pm_k$ is shorthand for $A^\pm(k_\perp, k_\parallel)$, $c_1$
461: and $c_2$ are dimensionless constants of order unity, and $S^\pm_k$
462: and $- \gamma^\pm_k A^\pm_k$ are forcing and damping terms,
463: respectively. The first term on the right-hand side of
464: equation~(\ref{eq:FPpm}) represents advection of fluctuation energy to
465: larger~$k_\perp$, while the second term represents diffusion of
466: fluctuation energy to larger~$|k_\parallel|$. The quantity
467: $\tau_{{\rm eff}, k_\perp}^\pm$ is an effective cascade time at
468: perpendicular scale~$k_\perp^{-1}$. Usually, the transfer of energy
469: to small scales is dominated by local interactions in $k$-space, and
470: the cascade time for a $w^+$ wave packet is $\sim (k_c^\mp
471: v_A)^{-1}$. In some cases, however, the shearing of small-scale wave
472: packets by much larger-scale wave packets can become important. To
473: account for such cases, the effective cascade time is taken to be
474: \begin{equation}
475: (\tau_{{\rm eff}, k_\perp}^\pm)^{-1}= \max\left[
476: \frac{q_\perp^4 A^\mp (q_\perp,0)}{v_A}\right]
477: \mbox{ \hspace{0.3cm} for $0<q_\perp< k_\perp$,}
478: \label{eq:taueff}
479: \end{equation}
480: i.e., $(\tau_{{\rm eff}, k_\perp}^\pm)^{-1}$ is the maximum value
481: of $k_c^\mp v_A$ for all perpendicular wave numbers between zero
482: and~$k_\perp$.
483: The flux of $w^\pm$ energy to larger~$k_\perp$ is
484: \begin{equation}
485: \epsilon^\pm(k_\perp) = 2\pi\int_{-\infty}^\infty dk_\parallel \,
486: \frac{c_1 k_\perp^2 A^\pm_k h^\pm_k}{\tau^\pm _{{\rm eff,} k_\perp}}.
487: \label{eq:epsilon}
488: \end{equation}
489: The term $h^\pm _k$ is given by
490: \begin{equation}
491: h^\pm _k = - \frac{1}{A^\pm(k_\perp,0)}\frac{\partial}{\partial k_\perp}
492: \left[k_\perp A^\pm (k_\perp,0)\right],
493: \label{eq:defh}
494: \end{equation}
495: and is included so that $\epsilon^\pm$ increases as the $A^\pm$
496: spectrum becomes a more steeply declining function of~$k_\perp$, in
497: accordance with weak turbulence theory (Galtier et al 2000, Lithwick
498: \& Goldreich~2003). To match the energy flux in weak turbulence
499: theory in the limit of zero cross helicity, one must set\footnote{The
500: value of~$c_1$ in equation~(\ref{eq:valc1}) is a factor of~2 larger
501: than the value that follows from the results of Galtier
502: et~al~(2000). It appears that this discrepancy results from the
503: omission of a factor of~2 in equation~(54) of Galtier
504: et~al~(2000). This can be seen by starting from equation~(46) of
505: Galtier et~al~(2000) and using the expression on page~1045 of Leith
506: \& Kraichnan~(1972) to simplify polar integrals of the form $\int
507: d^2 p\, d^2 q \,\delta (\bm{k} - \bm{p} - \bm{q})F(k,p,q)$ for
508: two-dimensional wave vectors $\bm{k}$, $\bm{p}$, and $\bm{q}$, where
509: $F$ is a function only of the wave-vector magnitudes and the
510: integral is over all values of $\bm{p}$ and $\bm{q}$.}
511: \begin{equation}
512: c_1 = -\frac{\pi J}{2},
513: \label{eq:valc1}
514: \end{equation}
515: where
516: \begin{equation}
517: J = \int_1^\infty dx\int_{-1}^1 dy\, \frac{
518: 2 [(x^2-1)(1-y^2)]^{1/2} (1+xy)^2 [8-(x+y)^3] \,\ln[(x+y)/2]}
519: {(x^2 - y^2)^4} \simeq -1.87.
520: \label{eq:defJ}
521: \end{equation}
522:
523:
524: For simplicity,
525: \begin{equation}
526: c_2 = 1.
527: \label{eq:valc2}
528: \end{equation}
529:
530:
531: \section{Steady-State Weak Turbulence}
532: \label{sec:weak}
533:
534: This section addresses weak turbulence in which
535: $\overline{ k}_\parallel^{\,+} \sim \overline{ k}_\parallel^{\,-}$
536: at the outer scale. The weak-turbulence condition, $\chi_{k_\perp}^\pm \ll 1$,
537: is equivalent to the condition $k_c^\pm
538: \ll \overline{ k}_\parallel^{\,\pm}$. Because $k_\parallel$-diffusion involves
539: a $\Delta k_\parallel \sim k_c^\pm$ during a time~$\tau_{k_\perp}^\pm$,
540: the $k_\parallel$-increment
541: over which energy diffuses
542: while cascading to larger~$k_\perp$ is much less than the
543: breadth of the spectrum in the $k_\parallel$ direction ($\sim
544: \overline{ k}_\parallel^{\,\pm}$),
545: so the $k_\parallel$-diffusion terms can be ignored to a good approximation.
546: In this case, equation~(\ref{eq:FPpm}) possesses
547: a steady-state solution in which $\epsilon^+$ and $\epsilon^-$ are constant,
548: and in which
549: \begin{equation}
550: A^\pm_k = g^\pm(k_\parallel) k_\perp^{-n^\pm},
551: \label{eq:ssw1}
552: \end{equation}
553: where $g^+ $ and $g^-$ are arbitrary functions of~$k_\parallel$, and where
554: \begin{equation}
555: n^+ + n^- = 6,
556: \label{eq:npnm}
557: \end{equation}
558: with $2<n^\pm < 4$. Equations~(\ref{eq:ssw1}) and (\ref{eq:npnm})
559: match the results of weak turbulence theory for incompressible MHD
560: turbulence if one allows only for three-wave interactions among shear
561: Alfv\'en waves (Galtier \& Chandran 2006), or if one considers only
562: the limit that $k_\perp \gg k_\parallel$ (Galtier et al 2002). If one
563: writes $n^\pm = 3 \pm \alpha$ with $|\alpha|<1$ and
564: sets $g^+(k_\parallel) = g^-(k_\parallel)$,
565: then equation~(\ref{eq:epsilon}) gives
566: \begin{equation}
567: \frac{\epsilon^+}{\epsilon^-} = \frac{2 + \alpha}{2-\alpha}.
568: \label{eq:fluxratio}
569: \end{equation}
570: In the limit~$\alpha \ll 1$, $\epsilon^+/\epsilon^- = 1+\alpha$,
571: in agreement with the weak-turbulence-theory result
572: for $k_\perp \gg |k_\parallel|$ (Lithwick \& Goldreich~2003),
573: as in the weak-turbulence
574: advection-diffusion model of Lithwick \& Goldreich~(2003).
575:
576: In steady state, $A^+(k_\perp,0)$ and $A^-(k_\perp,0)$ are forced to
577: be equal at the dissipation scale so that $\tau_k^+ = \tau_k^-$. This
578: phenomenon of ``pinning'' was discovered by Grappin et al~(1983) for
579: strong MHD turbulence, and further elaborated upon by Lithwick \&
580: Goldreich (2003) for the case of weak turbulence. The dominant
581: fluctuation type then has the steeper spectrum. If
582: $\epsilon^+/\epsilon^-$ is fixed, then the ratio $w_{k_f}^+/w_{k_f}^-$
583: of the rms amplitudes of the two fluctuation types at the outer
584: scale~$k_f^{-1}$ increases as $k_d/k_f$ increases, where $k_d$ is the
585: dissipation wave number. Alternatively, if $w_{k_f}^+/w_{k_f}^-$ is
586: fixed, then $\epsilon^+/\epsilon^-$ approaches unity as~$k_d/k_f
587: \rightarrow \infty$.
588:
589: Several of these results are illustrated by the numerical solution to
590: equation~(\ref{eq:FPpm}) shown in Figure~\ref{fig:f2}. This
591: solution is obtained using a logarithmic grid for $k_\perp$, with $k_{\perp,i} =
592: k_0 2^{i/n}$ for $0<i<N$. Similarly, $k_{\parallel,j} = k_0 2^{j/n}$
593: for $1<j<M$, but $k_{\parallel,j} = 0$ for $j=0$. $A^\pm_k$ is
594: advanced forward in time using a semi-implicit algorithm, in which the
595: terms $h_k^\pm$, $\tau_{{\rm eff},k_\perp}^\pm$, and $k_c^\pm$ on the
596: right-hand side of equation~(\ref{eq:FPpm}) are evaluated at the
597: beginning of the time step, and the $A^\pm$ terms on the right-hand
598: side of equation~(\ref{eq:FPpm}) are evaluated at the end of the time
599: step. The algorithm employs operator splitting, treating the $k_\perp$-advection,
600: forcing, and damping in one stage, and the $k_\parallel$-diffusion is
601: a second stage. In this approach, the matrix that has to be inverted
602: to execute each semi-implicit time step is tri-diagonal. An advantage
603: of this procedure over a fully explicit method is that the time step
604: is not limited by the $k_\parallel$-diffusion time at large~$k_\perp$
605: and small~$k_\parallel$. The discretized equations are written in
606: terms of the energy fluxes between neighboring cells, so that in the
607: absence of forcing and dissipation the algorithm conserves fluctuation
608: energy to machine accuracy. For the numerical solution plotted in
609: Figure~\ref{fig:f2}, $N=80$, $M=16$, $n=4$, $S^\pm = S_0^\pm k^2
610: \exp(- k^2/k_f^2)$, $S_0^+ = 1.2 S_0^-$, $k_f = 5 k_0$, and $\gamma^\pm_k
611: = 2k^2 \nu$, where $\nu$ is an effective viscosity. The
612: initial spectra are set equal to zero, and the equations are
613: integrated forward in time until a steady state is reached. In
614: steady state, $\delta v^+ = 2.5\times 10^{-3} v_A$ and $\delta
615: v^- = 6.4 \times 10^{-4} v_A$.
616:
617: \begin{figure}[h]
618: \includegraphics[width=2.1in]{f2a.eps}
619: \includegraphics[width=2.1in]{f2b.eps}
620: \includegraphics[width=2.1in]{f2c.eps}
621: \caption{\footnotesize
622: Numerical solution of equation~(\ref{eq:FPpm})
623: in the weak-turbulence limit.
624: {\em Left panel:} The dimensionless 1D
625: power spectrum defined in equation~(\ref{eq:defE}).
626: {\em Middle panel:} The spectral slopes at $k_\parallel=0$.
627: {\em Right panel:}
628: The weighted value of~$k_\parallel$ defined in equation~(\ref{eq:kparw}).
629: In all panels, the solid lines refer to~$w^+$ and the dashed
630: lines refer to~$w^-$. In the right-hand panel, the two lines
631: are almost on top of each other.
632: \label{fig:f2}}
633: \end{figure}
634:
635:
636: The left-hand panel of Figure~\ref{fig:f2} is a plot of the dimensionless
637: one-dimensional power spectrum,
638: \begin{equation}
639: E^\pm(k_\perp) = \frac{k_0 k_\perp}{v_A^2} \int_{-\infty}^\infty dk_\parallel \, A^\pm(k_\perp,k_\parallel),
640: \label{eq:defE}
641: \end{equation}
642: which is proportional to the energy per unit~$k_\perp$ in $w^\pm$
643: fluctuations. The middle panel of Figure~\ref{fig:f2}
644: shows that in the inertial range,
645: $d \ln A^+ (k_\perp,0)/d\ln k \simeq - 3.2$ and
646: $d \ln A^- (k_\perp,0)/d\ln k \simeq - 2.8$, as expected
647: for $S^+ /S^- = 1.2$. The right-hand panel shows that the
648: weighted value of $k_\parallel$,
649: \begin{equation}
650: \langle k_\parallel^\pm \rangle = \frac{\displaystyle
651: \int_{-\infty}^\infty dk_\parallel \, |k_\parallel|
652: A^\pm(k_\perp,k_\parallel)}{\displaystyle \int_{-\infty}^\infty
653: dk_\parallel \, A^\pm(k_\perp,k_\parallel)},
654: \label{eq:kparw}
655: \end{equation}
656: is roughly constant in the inertial range.
657:
658: \section{Steady-State Strong Turbulence}
659: \label{sec:strong}
660:
661: This section addresses strong turbulence with $\chi_{k_f}^+ \sim 1$,
662: $\chi_{k_f}^- \la 1$, $w_{k_f}^+ \geq w_{k_f}^-$, and $\overline{
663: k}_\parallel^{\,+} \sim \overline{ k}_\parallel^{\,-} $ at the outer
664: scale~$k_f^{-1}$. The discussion allows for the possibility that
665: $\chi_{k_\perp}^- \ll 1$. As fluctuation energy cascades to
666: larger~$k_\perp$, it diffuses to larger~$|k_\parallel|$, so that
667: $\overline{ k}_\parallel^{\,+}$ and $\overline{ k}_\parallel^{\,-}$
668: increase with increasing~$k_\perp$. Moreover, for both $w^+$ and
669: $w^-$, the fluctuation energy diffuses over a $k_\parallel$-increment
670: of $\sim k_c^+$ during one cascade time. For the steady-state
671: solutions of interest, $k_c^+$ is an increasing function of~$k_\perp$,
672: and thus at each $k_\perp$ we will have that $\overline{
673: k}_\parallel^{\,+} \sim \overline{ k}_\parallel^{\,-} \sim
674: k_c^+$. One can thus define a single parallel-wavenumber, $\overline{
675: k_\parallel}(k_\perp)$, to describe the spectra, with
676: \begin{equation}
677: \overline{ k_\parallel} \sim
678: \overline{ k}_\parallel^{\,+} \sim \overline{ k}_\parallel^{\,-}
679: \sim k_c^+
680: \label{eq:defolkp}
681: \end{equation}
682: at each~$k_\perp$. Since $\overline{ k}_\parallel^{\,+}
683: \sim k_c^+$ at each scale,
684: \begin{equation}
685: \chi_{k_\perp}^+ \sim 1
686: \label{eq:chistrong}
687: \end{equation}
688: throughout the inertial range.
689: On the other hand, since $w_{k_\perp}^-$ can be much
690: less than~$w_{k_\perp}^+$, $\chi_{k_\perp}^-$ can be~$\ll 1$.
691:
692: The cascade time for the $w^-$ fluctuations is given by the strong-turbulence
693: phenomenology of equation~(\ref{eq:tau2}), so that the energy
694: flux in $w^-$ fluctuations is
695: \begin{equation}
696: \epsilon^- \sim
697: \frac{(w^-_{k_\perp})^2}{\tau_{k_\perp}^-}
698: \sim k_\perp w_{k_\perp}^+ (w_{k_\perp}^-)^2.
699: \label{eq:epsms}
700: \end{equation}
701: Allowing for the possibility that $\chi_{k_\perp}^- \ll 1$, the cascade
702: time of the $w^+$ fluctuations follows the weak-turbulence
703: phenomenology of equation~(\ref{eq:tau1}). This formula
704: is also accurate for $\chi_{k_\perp}^-$ as large as~1 (in which
705: case $w^-_{k_\perp} \sim w^+_{k_\perp}$). The energy
706: flux in $w^+$ fluctuations is then
707: \begin{equation}
708: \epsilon^+ \sim \frac{(w^+_{k_\perp})^2}{\tau_{k_\perp}^+}
709: \sim \frac{k_\perp^2 (w_{k_\perp}^-)^2 (w_{k_\perp}^+)^2 }{\overline{ k_\parallel} v_A}
710: \sim k_\perp w_{k_\perp}^+ (w_{k_\perp}^-)^2,
711: \label{eq:epsps}
712: \end{equation}
713: which is roughly the same as~$\epsilon^-$. It is assumed that the
714: energy flux depends on the spectral slope as in weak turbulence, so
715: that the fluctuation type with the steeper spectrum has the larger
716: energy flux. If
717: \begin{equation}
718: w^\pm \propto k_\perp^{-a^\pm},
719: \label{eq:defapam}
720: \end{equation}
721: then equations~(\ref{eq:epsms}) and (\ref{eq:epsps}) imply that
722: when $\epsilon^+$ and $\epsilon^-$ are independent of~$k_\perp$,
723: \begin{equation}
724: a^+ + 2a^- = 1.
725: \label{eq:apam1}
726: \end{equation}
727: The condition that $\chi_{k_\perp}^+ \sim 1$ throughout the inertial range
728: then implies that
729: \begin{equation}
730: \overline{ k_\parallel} \propto k_\perp^{1- a^+}.
731: \label{eq:olkp1}
732: \end{equation}
733: As discussed in earlier studies (Grappin et~al~1983, Lithwick \&
734: Goldreich~2003), the spectra are pinned at the dissipation scale, so
735: that the dominant fluctuation type will have the steeper spectrum and
736: a somewhat larger energy flux. For the zero-cross-helicity case,
737: equations~(\ref{eq:apam1}) and (\ref{eq:olkp1}) give $w_{k_\perp}^+ =
738: w_{k_\perp}^- \propto k_\perp^{-1/3}$ and $ \overline{ k}_\parallel
739: \propto k_\perp^{2/3}$, as in the work of Goldreich \& Sridhar (1995)
740: [see also Higdon (1984)].
741:
742:
743: When $S_k^\pm = \gamma_k^\pm = 0$,
744: equation~(\ref{eq:FPpm}) possesses
745: an analytical solution that reproduces the above scalings.
746: This solution can be obtained by starting with the assumptions that
747: \begin{equation}
748: A^\pm (k_\perp,0) = c_3^\pm k_\perp^{-b^\pm},
749: \label{eq:ansatz}
750: \end{equation}
751: that the energy cascade is
752: dominated by local interactions, and that
753: $A^+(k_\perp,0) > A^-(k_\perp,0)$
754: for all~$k_\perp$. Equation~(\ref{eq:taueff}) then
755: becomes $(\tau_{{\rm eff},k_\perp}^\pm)^{-1} = k_\perp^4
756: A^\mp(k_\perp,0)/v_A$, and $k_{c,\rm max} = k_c^+$. Upon defining
757: \begin{equation}
758: f^\pm_k = k_\perp^{6 - b^\mp}A_{k}^\pm
759: \label{eq:deff}
760: \end{equation}
761: and
762: \begin{equation}
763: s = k_\perp^{8-2b^+},
764: \label{eq:defs}
765: \end{equation}
766: one can rewrite equation~(\ref{eq:FPpm}) as
767: \begin{equation}
768: \frac{\partial f^\pm_k }{\partial s} = D^\pm \frac{\partial ^2 f_k^\pm}{
769: \partial k_\parallel^2},
770: \label{eq:eqnf}
771: \end{equation}
772: with
773: \begin{equation}
774: D^\pm = \frac{c_2 (c_3^+)^2}{c_1 (8-2b^+)(b^\pm -1) v_A^4}.
775: \label{eq:defDpm}
776: \end{equation}
777: Equation~(\ref{eq:eqnf}) is solved by taking
778: \begin{equation}
779: f_k^\pm = \frac{c_4^\pm}{\sqrt{s}}\exp\left(- \frac{k_\parallel^2}{4D^\pm s}\right).
780: \label{eq:fsolve}
781: \end{equation}
782: Requiring that equation~(\ref{eq:ansatz})
783: be satisfied, one finds that $c_4^\pm = c_3^\pm$
784: and
785: \begin{equation}
786: 2b^+ + b^- = 10.
787: \label{eq:bpbm}
788: \end{equation}
789: The dominance of local interactions requires that $b^+ < 4$,
790: and thus $b^- > 2$. When forcing and dissipation are taken into account,
791: the exact solution becomes an approximate solution that is valid
792: only within the inertial range. In this case, $b^+ > b^-$
793: because the spectra are pinned at the dissipation
794: scale whereas $A^+(k_\perp,0)$ is larger
795: than~$A^-(k_\perp,0)$ within the inertial range.
796: Equation~(\ref{eq:fsolve}) implies that
797: \begin{equation}
798: \overline{ k}_\parallel^{\,+} \simeq
799: \overline{ k}_\parallel^{\,-} \sim \sqrt{D^\pm s} \sim \frac{c_3^+ k_\perp^{4-b^+}}{v_A^2},
800: \label{eq:kpkm0}
801: \end{equation}
802: where the dimensionless constants in the
803: expression for~$D^\pm$ have been dropped, but
804: $c_3^+$, which has dimensions, has been kept.
805: Equations~(\ref{eq:defkc}), (\ref{eq:ansatz}), and~(\ref{eq:kpkm0})
806: show that $k_c^+ \sim \overline{ k}_\parallel^{\,+}$ for all $k_\perp$, so
807: that $\chi_{k_\perp}^+ \sim 1$ for all~$k_\perp$.
808: Equation~(\ref{eq:defE}) gives
809: \begin{equation}
810: E^\pm(k_\perp) \propto A^\pm(k_\perp,0) k_\perp \overline{ k}_\parallel^{\,+} \propto
811: k_\perp^{5-b^+ - b^\pm},
812: \label{eq:Eb}
813: \end{equation}
814: from which it follows that
815: \begin{equation}
816: w_{k_\perp}^+ \propto k_\perp^{3-b^+},
817: \label{eq:wlps}
818: \end{equation}
819: and
820: \begin{equation}
821: w_{k_\perp}^- \propto k_\perp^{(6- b^+ - b^-)/2}.
822: \label{eq:wlms}
823: \end{equation}
824: This solution reduces to the critical-balance solution of Goldreich \&
825: Sridhar (1995) when $b^+ = b^- = 10/3$, in which case $\overline{
826: k}_\parallel^{\, \pm} \propto k_\perp^{2/3}$, and $w_{k_\perp}^\pm \propto
827: k_\perp^{-1/3}$. Comparing equations~(\ref{eq:wlps}) and
828: (\ref{eq:wlms}) with equation~(\ref{eq:defapam}), it can be seen that $b^+$
829: corresponds to $3+a^+$ and $b^-$ corresponds to $3 + 2a^- - a^+$.
830: Equation~(\ref{eq:kpkm0}) is thus equivalent to
831: equation~(\ref{eq:olkp1}), and equation~(\ref{eq:bpbm}) is equivalent
832: to equation~(\ref{eq:apam1}).
833:
834: Figure~\ref{fig:f3} shows the results from a numerical solution of
835: equation~(\ref{eq:FPpm}), obtained by integrating forward in time as
836: described in section~\ref{sec:weak} with the spectra initially equal
837: to zero. The numerical solution was obtained by setting $S^\pm = S_0^\pm
838: k^2 \exp(- k^2/k_f^2)$ with $S_0^+ =
839: 1.2 S_0^-$ and using the parameters
840: (defined in section~\ref{sec:weak}) $N=80$, $M=60$,
841: $n=4$, $k_f= 5 k_0$, and $\gamma_k^\pm = 2k^2 \nu$, where the constant $\nu$ is
842: an effective viscosity. The rms velocities at steady
843: state are $\delta v^+ = 1.5 v_A$ and $\delta v^- = 0.10 v_A$. The
844: left-hand panel shows that the one-dimensional energy
845: spectrum~$E^+_{k_\perp}$ is $\propto k_\perp^{-2.14}$ in the inertial
846: range, which corresponds to~$b^+ = 3.57$ in equation~(\ref{eq:Eb}).
847: Equation~(\ref{eq:bpbm}) then gives $b^- = 2.86$. The dotted lines in
848: the middle panel of Figure~\ref{fig:f3} correspond to the values of
849: $b^+ = 3.57$ and $b^- = 2.86$, which are reasonably close to the
850: values of $- d \ln A^\pm(k_\perp,0)/d\ln k_\perp$ in the numerical
851: solution, although these latter values vary throughout the inertial
852: range in the numerical solution. For $b^+ = 3.57$,
853: equation~(\ref{eq:kpkm0}) gives $\overline{ k}_\parallel^{\,+}\propto
854: k_\perp^{0.43}$, which is a close match to the numerical solution, as
855: shown in the right-hand panel of Figure~\ref{fig:f3}. The left-hand
856: panel of Figure~\ref{fig:f3} shows that the steady-state solutions for
857: $A^+$ and $A^-$ are ``pinned'' at the dissipation scale, as expected.
858:
859: \begin{figure}[h]
860: \includegraphics[width=2.1in]{f3a.eps}
861: \includegraphics[width=2.1in]{f3b.eps}
862: \includegraphics[width=2.1in]{f3c.eps}
863: \caption{\footnotesize Numerical solution of equation~(\ref{eq:FPpm})
864: for strong turbulence with~$\chi_{k_\perp}^+\sim 1$.
865: {\em Left panel:} The dimensionless 1D
866: power spectrum defined in equation~(\ref{eq:defE}).
867: {\em Middle panel:} The spectral slopes at $k_\parallel=0$.
868: {\em Right panel:}
869: The weighted value of~$k_\parallel$ defined in equation~(\ref{eq:kparw}).
870: In all panels, the solid lines refer to~$w^+$ and the dashed
871: lines refer to~$w^-$.
872: \label{fig:f3}}
873: \end{figure}
874:
875: It should be noted that when $\chi_{k_\perp}^+ \sim 1$ and
876: $\chi_{k_\perp}^- \ll 1$, the dominant~$w^+$ fluctuations are only
877: weakly damped by nonlinear interactions with $w^-$ waves, in the sense
878: that $\tau_{k_\perp}^+$ is much larger than the linear wave period.
879: On the other hand, for the smaller-amplitude $w^-$
880: fluctuations, the linear wave period and cascade time are comparable.
881: Thus, paradoxically, the larger-amplitude~$w^+$ fluctuations
882: can be described as waves, or, more precisely, a non-sinusoidal
883: wave train, whereas the smaller-amplitude $w^-$ fluctuations
884: can not be accurately described as waves.
885:
886: \section{Transition Between Weak Turbulence and Strong Turbulence}
887: \label{sec:trans}
888:
889: This section again addresses turbulence in which
890: $\overline{k}_\parallel^{\,+} \sim\overline{k}_\parallel^{\,-} $ at
891: the outer-scale wavenumber,~$k_f$. In the weak-turbulence limit,
892: $\chi_{k_\perp}^+$ and~$\chi_{k_\perp}^-$ increase with
893: increasing~$k_\perp$. If the dissipation wavenumber~$k_d$ is
894: sufficiently large, then $\chi_{k_\perp}^+$ and/or $\chi_{k_\perp}^-$
895: will increase to a value of order unity at some $k_\perp$ within the
896: inertial range. This perpendicular wavenumber is denoted~$k_{\rm
897: trans}$. The turbulence will then be described by the
898: weak-turbulence scalings of section~\ref{sec:weak} for~$k_f\ll k_\perp
899: \ll k_{\rm trans}$, and by the strong-turbulence scalings of
900: section~\ref{sec:strong} for $k_{\rm trans} \ll k_\perp \ll k_d$.
901: Figure~\ref{fig:f4} shows a numerical solution of
902: equation~(\ref{eq:FPpm}) that illustrates how the turbulence makes
903: this transition in a smooth manner. At small wavenumbers, this
904: solution is similar to the weak-turbulence solution plotted in
905: Figure~\ref{fig:f2}, and at large wavenumbers it is similar to the
906: strong-turbulence solution plotted in Figure~\ref{fig:f3}. The
907: solution shown in Figure~\ref{fig:f4} was obtained by integrating
908: equation~(\ref{eq:FPpm}) forward in time to steady state using the
909: numerical method described in section~\ref{sec:weak}. The spectra were
910: initially set equal to zero. The numerical solution was obtained by
911: setting $S^\pm = S_0^\pm k^2 \exp(- k^2/k_f^2)$ with $S_0^+ = 1.2
912: S_0^-$ and using the parameters $N=100$, $M=56$, $n=4$, $k_f= 5 k_0$,
913: and $\gamma_k^\pm = 2k^2 \nu$, where the constant $\nu$ is an
914: effective viscosity. The rms velocities at steady state are $\delta
915: v^+ = 0.32 v_A$ and $\delta v^- = 0.012 v_A$.
916:
917: \begin{figure}[h]
918: \includegraphics[width=2.1in]{f4a.eps}
919: \includegraphics[width=2.1in]{f4b.eps}
920: \includegraphics[width=2.1in]{f4c.eps}
921: \caption{\footnotesize Numerical solution of equation~(\ref{eq:FPpm})
922: showing a smooth transition from the
923: weak-turbulence limit at small~$k_\perp$ to the strong-turbulence
924: limit at large~$k_\perp$.
925: {\em Left panel:} The dimensionless 1D
926: power spectrum defined in equation~(\ref{eq:defE}).
927: {\em Middle panel:} The spectral slopes at $k_\parallel=0$.
928: {\em Right panel:}
929: The weighted value of~$k_\parallel$ defined in equation~(\ref{eq:kparw}).
930: In all panels, the solid lines refer to~$w^+$ and the dashed
931: lines refer to~$w^-$.
932: \label{fig:f4}}
933: \end{figure}
934:
935: \section{Unequal Parallel Correlation Lengths at the Outer Scale}
936: \label{sec:unequal}
937:
938: In sections~\ref{sec:weak} through \ref{sec:trans}, it was assumed
939: that $\overline{ k}_\parallel^{\, + } \sim \overline{ k}_\parallel^{\,
940: - }$ at the outer scale. This assumption is applicable to many
941: settings. For example, in a plasma of dimension~$L$ that is stirred by
942: a force that has a correlation length~$l \ll L$, the velocity
943: fluctuations that are excited have a correlation length~$l$, and this
944: correlation length is imprinted on both the~$w^+$ and $w^-$
945: fluctuations. On the other hand, if waves are launched along the
946: magnetic field into a bounded plasma from opposite sides of the
947: plasma, and the waves from one side have a much larger parallel
948: correlation length than the waves from the other side, it is
949: possible to set up turbulence in which the two wave types have very
950: different parallel correlation lengths at the outer scale. This
951: situation is discussed briefly in this section.
952:
953: For strong turbulence, if both $\chi_{k_\perp}^+$ and
954: $\chi_{k_\perp}^-$ are $\sim 1$ at some perpendicular
955: scale~$k_\perp^{-1}$, but one fluctuation type, say $w^+$, has a much
956: smaller parallel correlation length than the other (and thus a much
957: larger amplitude), then during a time~$\tau_{k_\perp}^-$ the
958: ``propagation with distortion'' mechanism discussed in
959: section~\ref{sec:parcasc} will increase $\overline{
960: k}_\parallel^{\,-}$ until it equals $\overline{ k}_\parallel^{\,+}$,
961: which will cause $\chi_{k_\perp}^-$ to become~$\ll 1$ at
962: scale~$k_\perp^{-1}$. At smaller scales, the solution can be
963: described by the scalings presented in section~\ref{sec:strong}, in
964: which $\overline{ k}_\parallel^{\,+}(k_\perp) \sim \overline{
965: k}_\parallel^{\,-}(k_\perp)$. Similarly, if $\chi_{k_\perp}^+ \sim
966: 1$ but $\chi_{k_\perp}^- \ll 1$ at some scale $k_\perp^{-1}$ and if
967: $\overline{ k}_\parallel^{\,+} \gg \overline{ k}_\parallel^{\,-}$ at
968: that scale, then during a time~$\tau_{k_\perp}^-$ the ``propagation
969: with distortion'' mechanism discussed in section~\ref{sec:parcasc}
970: will again increase $\overline{ k}_\parallel^{\,-}$ until it equals
971: $\overline{ k}_\parallel^{\,+}$, the parallel scales will remain
972: comparable at smaller perpendicular scales, and the solution can be
973: described by the scalings in section~\ref{sec:strong}. The case in
974: which $ \chi_{k_\perp}^+ \sim 1$, $\chi_{k_\perp}^- \ll 1$, and
975: $\overline{ k}_\parallel^{\,+} \ll \overline{ k} _\parallel^{\,-}$ is
976: not addressed in this paper.
977:
978:
979: \section{Implications for Turbulence in the Solar Corona and Solar Wind}
980: \label{sec:solarwind}
981:
982: In this section, the preceding analysis of incompressible MHD
983: turbulence is applied to the solar wind and solar corona. It should be
984: noted at the outset, however, that the solar wind and solar corona
985: (beyond roughly~$r=1.5R_{\sun}$, where $r$ is distance from the Sun's
986: center) are in the collisionless regime, and the pressure tensor is
987: not isotropic as assumed in ideal MHD. Moreover, the value of $\beta =
988: 8\pi p/B^2$ is $\ll 1$ in the corona and typically $\sim 1$ in the
989: solar wind at 1~AU, whereas incompressible MHD corresponds to the
990: limit~$\beta \rightarrow \infty$. A preliminary question that needs to
991: be addressed is thus the extent to which incompressible MHD is an
992: accurate model for these plasmas.
993:
994: Schekochihin et~al~(2007) have recently carried out extensive
995: calculations based on kinetic theory that provide a detailed answer to
996: this question. These authors examined anisotropic turbulence in weakly
997: collisional magnetized plasmas using gyrokinetics, a low-frequency
998: expansion of the Vlasov equation that averages over the gyromotion of
999: the particles. By applying the form of the gyrokinetic expansion
1000: derived by Howes et~al~(2006), Schekochihin et~al~(2007) showed
1001: analytically that non-compressive Alfv\'enic turbulence in the
1002: quasi-2D regime (i.e., $k_\perp \gg k_\parallel$) can be accurately
1003: described using reduced MHD in both the collisional and collisionless
1004: limits, regardless of~$\beta$, provided that the length scales of the
1005: fluctuations are much larger than the proton gyroradius and the
1006: frequencies are much less than the proton cyclotron frequency. Since
1007: non-compressive quasi-2D fluctuations are thought to be the dominant
1008: component of the turbulence in the solar wind (see, e.g., Bieber
1009: et~al~1994) and the solar corona (Dmitruk \& Matthaeus~2003, Cranmer
1010: \& van Ballegooijen~2005), incompressible MHD is a useful
1011: approximation for modeling turbulence in these settings.
1012:
1013: \subsection{Cross helicity in the solar wind and solar corona}
1014: \label{sec:CH}
1015:
1016: Cross helicity in the solar wind has been measured {\em in situ} by
1017: several different spacecraft. In terms of the Elsasser variables
1018: $w^\pm$, there is a substantial excess of outward propagating
1019: fluctuations (taken to be $w^+$ throughout this section) over inward
1020: propagating fluctuations (taken to be $w^-$) in the inner heliosphere,
1021: although this imbalance decreases with increasing~$r$, as seen, for
1022: example, in Voyager data for low heliographic latitude (Matthaeus \&
1023: Goldstein 1982, Roberts et~al~1987) and Ulysses data at high latitude
1024: (Goldstein et~al~1995). In a study of Ulysses and Helios data,
1025: Bavassano et~al~(2000) found that $e^+/e^- \propto r^{-1.02}$ for
1026: $r<2.6$~AU and $e^+/e^- \sim 2$ for $3\mbox{ AU}\la r \la 5 \mbox{
1027: AU}$, where $e^\pm$ is the energy per unit mass associated with
1028: $w^\pm$ fluctuations. These numbers are intended as illustrative
1029: average values, as individual measurements of $e^+/e^-$ in the solar
1030: wind vary significantly.
1031:
1032: Although it has not been directly measured, the ratio $e^+/e^-$ is
1033: likely very large in open-field-line regions of the solar corona.
1034: This can be seen from the work of Cranmer \& van Ballegooijen (2005),
1035: who modeled the generation of Alfv\'en waves by the observed motions
1036: of field-line footpoints in the photosphere, and the propagation and
1037: reflection of these waves as they travel along open field lines from
1038: the photosphere out into the interplanetary medium. They found that
1039: the ratio of the frequency-integrated rms Elsasser variables ($w^+$
1040: and $w^-$) is $\sim 30$ at $r= 2R_{\sun}$ (i.e., $e^+/e^- \sim 900$).
1041: Verdini \& Velli~(2007) developed a different model for the
1042: propagation, reflection, and turbulent dissipation of Alfv\'en waves
1043: in the solar atmosphere and solar wind and found that $e^+/e^-\simeq
1044: 80$ at $r=2R_{\sun}$. Based on these results, one can make the rough
1045: estimate that
1046: \begin{equation}
1047: \frac{w_{k_f}^+}{w_{k_f}^-} \sim 10 \hspace{1cm} \mbox{ (at $r=2R_{\sun}$),}
1048: \label{eq:wpwmra}
1049: \end{equation}
1050: where $k_f$ is the perpendicular wavenumber at the outer scale.
1051:
1052:
1053: \subsection{Is quasi-2D turbulence in the corona and
1054: solar wind weak or strong?}
1055: \label{sec:q2dws}
1056:
1057: In much of the solar wind, $\delta B$ is comparable to $B_0$, and the
1058: turbulence is in the strong-turbulence regime with~$\chi_{k_f}^+ \sim
1059: 1$. For the corona, Cranmer \& van Ballegooijen (2005) found that the
1060: outer-scale fluctuations in open-field-line regions have periods of
1061: $T= 1-5$ minutes, $\delta v \sim 100$~km/s, and perpendicular
1062: correlation lengths of~$L_\perp \sim k_f^{-1} \sim 10^4$~km. The
1063: Alfv\'en speed in their model corona is between 2000 and 3000~km/s at
1064: $r=2R_{\sun}$, and thus $\delta v \ll v_A$. However, the parallel
1065: correlation length $L_\parallel$ of the outer-scale fluctuations is
1066: $\sim v_A T = 1.8 -9 \times 10^5$~km, which is $\gg L_\perp$. Because
1067: $L_\parallel/L_\perp \sim v_A/\delta v$,
1068: \begin{equation}
1069: \chi_{k_f}^+ \sim 1
1070: \label{eq:chicorona}
1071: \end{equation}
1072: and the low-frequency fluctuations launched into
1073: the corona by footpoint motions are in the strong-turbulence regime.
1074: There may be an additional population of higher-frequency waves in the
1075: weak-turbulence regime, but these are not discussed here.
1076:
1077:
1078: \subsection{Parallel correlation lengths of inward and outward waves}
1079: \label{sec:pcl}
1080:
1081: In open-field-line regions of the corona, when $w^+$ waves are
1082: reflected, the resulting $w^-$ waves have the same frequencies as the
1083: $w^+$ waves. On the other hand, wave-reflection is more efficient at
1084: lower frequencies (Velli 1993), so if there is a range of wave
1085: frequencies at each~$k_\perp$, the energy-weighted average frequency
1086: of inward waves would tend to be somewhat lower than that of the
1087: outward waves. This suggests that at the outer scale the parallel
1088: correlation length~$L_\parallel$ of the $w^-$ fluctuations is somewhat
1089: larger than the value of~$L_\parallel$ for the $w^+$ fluctuations.
1090: However, given equation~(\ref{eq:chicorona}), the $w^+$ fluctuations
1091: imprint their parallel correlation length on the $w^-$ fluctuations
1092: during a single turnover time $\tau_{k_\perp}^-$, as argued in
1093: section~\ref{sec:unequal}. The parallel correlation lengths of the
1094: $w^+$ and $w^-$ fluctuations in the corona can thus be taken to be
1095: approximately equal at the outer scale, and hence also at smaller
1096: scales. The same approximation is reasonable for turbulence in the
1097: solar wind.
1098:
1099: \subsection{Energy Dissipation Rate}
1100: \label{sec:epsilon}
1101:
1102: If we take $k_f$ to be the perpendicular wave number at the outer
1103: scale, $w_{k_f}^+$ to be the rms amplitude of the outward-propagating
1104: fluctuations at the outer scale, and $w_{k_f}^-$ to be the rms
1105: amplitude of the Sunward-propagating fluctuations at the outer scale,
1106: then equations~(\ref{eq:epsms}), (\ref{eq:epsps}),
1107: and~(\ref{eq:chicorona}) imply that
1108: \begin{equation}
1109: \epsilon^+ \sim \epsilon^- \sim k_f w^+_{k_f} (w^-_{k_f})^2.
1110: \label{eq:epsSW}
1111: \end{equation}
1112: This estimate is a factor of $\sim w_{k_f}^-/w_{k_f}^+$ smaller than
1113: the standard strong-turbulence estimate of $\epsilon^+ \sim k_f
1114: (w_{k_f}^+)^2 w_{k_f}^-$ that appears in many studies (e.g., Zhou \&
1115: Matthaeus 1990, Cranmer \& Van Ballegooijen~2005, Lithwick, Goldreich,
1116: \& Sridhar~2007, Verdini \& Velli~2007). This difference has important
1117: implications for turbulent heating of the solar corona and solar wind.
1118:
1119: \subsection{Cascade Time}
1120: \label{sec:cascadetime}
1121:
1122:
1123: For the energetically dominant $w^+$ fluctuations, the cascade
1124: time~$\tau_{k_\perp}^+$ is much longer than the linear wave period, at
1125: least at scales much larger than the dissipation scale. This result
1126: is important for determining the conditions under which turbulence can
1127: be a viable mechanism to explain the heating of the solar
1128: corona. Observations taken with the {\em Ultraviolet Coronagraph
1129: Spectrometer} (UVCS) indicate that there is strong heating of
1130: coronal plasma at $r\lesssim 2 R_{\sun}$ (Kohl et al 1998, Antonucci
1131: et al 2000). An appealing model to explain this heating is that
1132: low-frequency Alfv\'en waves are launched by turbulent motions of
1133: field-line footpoints in the photosphere, that some of these waves are
1134: reflected, and that interactions between oppositely directed Alfv\'en
1135: wave packets in the corona causes the wave energy to cascade to small
1136: scales and dissipate (Matthaeus et~al~1999, 2002; Dmitruk et~al~2001,
1137: 2002; Cranmer \& van Ballegooijen 2005, 2007; Verdini \& Velli 2007).
1138: In one version of this model, the waves that cross the transition
1139: region and enter the corona have not yet undergone a turbulent
1140: cascade, and their energy is concentrated at the fairly long periods
1141: ($>1$~minute) characteristic of the observed footpoint motions
1142: that are believed to make the dominant contribution to the outward
1143: directed wave flux. In order for this scenario to explain the UVCS
1144: measurements, there needs to be time for the outer-scale waves to
1145: cascade within the corona before they travel beyond~$r\simeq 2
1146: R_{\sun}$. If, as above, we take $k_f$ to be the value of~$k_\perp$
1147: at the outer scale, $w_{k_f}^+$ to be the rms amplitude of the outward
1148: waves at the outer scale, and $L_\parallel$ to be the parallel
1149: correlation length of the fluctuations at the outer scale, then
1150: equation~(\ref{eq:tau1}) can be used to express the cascade time for
1151: the outward waves at the outer scale as
1152: \begin{equation}
1153: \tau^+_{k_f} \sim \frac{L_\parallel}{v_A} \left(\frac{w_{k_f}^+}{w_{k_f}^-}\right)^2
1154: (\chi_{k_f}^+)^{-2},
1155: \label{eq:taucorona}
1156: \end{equation}
1157: where $\chi_{k_f}^+ \sim k_f w_{k_f}^+ L_\parallel/v_A$. Thus, for
1158: waves with a period $L_\parallel/v_A \sim 1$~minute,
1159: equations~(\ref{eq:wpwmra}), (\ref{eq:chicorona}),
1160: and~(\ref{eq:taucorona}) give $\tau^+_{k_f} \sim 100$~minutes. On the
1161: other hand, the Alfv\'en speed in a coronal hole at~$r< 2R_{\sun}$
1162: is~$\sim 2000-3000$~km/s (Cranmer \& van Ballegooijen 2005), and the
1163: time for an Alfv\'en wave to travel from the coronal base out to
1164: $r=2R_{\sun}$ is 4-6 minutes. There is thus not enough time for the
1165: energy of waves with periods $> 1$~minute to cascade and dissipate
1166: within a few solar radii of the Sun.
1167:
1168: Dmitruk \& Matthaeus~(2003) and Verdini \& Velli (2007) avoid this
1169: difficulty by postulating that a broad frequency spectrum of waves is
1170: launched upwards from the photosphere. Another possible way around
1171: this difficulty is the development of a broad frequency spectrum of
1172: fluctuations from wave-packet collisions in the chromosphere, in which
1173: the energies of inward and outward propagating waves are
1174: comparable due to strong wave reflection at the transition region.
1175:
1176:
1177:
1178: \subsection{Spectral Index}
1179: \label{sec:slope}
1180:
1181: Much of the discussion of the inertial-range power spectrum of
1182: solar-wind turbulence has focused on the question of whether the
1183: spectral index is closer to the Kolmogorov (1941) value of~$-5/3$ or
1184: the Iroshnikov-Kraichnan value of~$-3/2$ (Iroshnikov~1963,
1185: Kraichnan~1965). A value of $-5/3$ is supported by a number of
1186: theoretical studies (e.g., Montgomery \& Turner 1981, Higdon 1984,
1187: Goldreich \& Sridhar 1995) and numerical simulations (Cho \& Vishniac
1188: 2000, M\"uller \& Biskamp 2000, Cho et~al~ 2002, Haugen et~al~2004). A
1189: value value of~$-3/2$ is supported by a second group of theoretical
1190: studies (Boldyrev 2005, 2006; Mason et~al~2006; see also Beresnyak \&
1191: Lazarian 2006) and numerical simulations (Maron \& Goldreich 2001,
1192: M\"uller et~al~2003, M\"uller \& Grappin 2005, Mininni \&
1193: Pouquet~2007). It should be noted that all of the above-mentioned
1194: studies address MHD turbulence with negligible cross helicity.
1195:
1196: Spacecraft measurements yield frequency spectra for the magnetic field
1197: and velocity fluctuations, where the frequency $f$ is approximately
1198: $k_r U/2\pi$, where $k_r$ is the radial component of the wave-vector
1199: and $U$ is the solar-wind speed [Taylor's (1938) ``frozen-in flow
1200: hypothesis'']. Below a spectral-break frequency~$f_b$, the spectra
1201: are typically fairly flat, being approximately proportional
1202: to~$f^{-1}$ (Matthaeus \& Goldstein 1986). At $f>f_b$, the spectra
1203: steepen. The time scale corresponding to the spectral break,
1204: $f_b^{-1}$, increases with increasing~$r$. For example, Bruno \&
1205: Carbone (2005) found that $f_b^{-1}$ was $0.06$~hours at 0.3~AU,
1206: 0.16~hours at 0.7~AU, and 0.4~hours at 0.9~AU in a sample of Helios~2
1207: data. In two other studies based on data from several spacecraft,
1208: Matthaeus \& Goldstein (1986) found that $f_b^{-1}\sim 3.5$~hours at
1209: 1~AU, while Klein et~al~(1992) found $f_b^{-1} \sim 12$~hours at
1210: 4~AU. The inertial range roughly corresponds to frequencies in the
1211: interval $f_b < f < f_d$, where~$f_d$ is the frequency corresponding
1212: to the dissipation scale. At 1~AU, $f_d \sim 0.3 \mbox{ s}^{-1}$
1213: (Smith et~al~2006). A large number of inertial-range spectral indices
1214: have been reported in the literature. For example, Matthaeus \&
1215: Goldstein (1982) found a spectral index of $-1.73 \pm 0.08$ for the
1216: magnetic field in Voyager data at $r=1$~AU, and a spectral index of
1217: $-1.69 \pm 0.08$ for the total energy. Goldstein et~al~(1995, Fig. 1)
1218: found that the spectral index for the $w^+$ fluctuations was slightly
1219: steeper than $-5/3$ in Ulysses data at 2~AU and 4~AU. Their results
1220: also suggest a shallower $w^-$ spectrum, consistent with the idea that
1221: the spectra are pinned at the dissipation wavenumber~$k_d$. In a
1222: study of Helios~2 magnetic-field data, Bruno \& Carbone (2005, Figure
1223: 23) found spectral indices of $-1.72$ at 0.3~AU, $-1.67$ at 0.7~AU,
1224: and $-1.70$ at 0.9~AU. Using data from the WIND spacecraft at 1~AU,
1225: Podesta et~al~(2007) found a total-energy spectral index of $-1.63$,
1226: with the velocity spectrum flatter than the magnetic spectrum. Marsch
1227: \& Tu (1996) found a spectral index of $-1.65\pm 0.01$ for the
1228: magnetic field at 1~AU in Helios~2 data. Horbury \& Balogh (1995)
1229: found a spectral index close to~$-5/3$ for the magnetic field in
1230: Ulysses data at 2.5~AU. Using magnetic-field data from the ACE
1231: spacecraft at 1~AU, Smith et~al~(2006) found a spectral index of
1232: $-1.63\pm 0.14$ in open-field-line regions and $-1.56\pm 0.16$ in
1233: magnetic clouds. Smith (2003) found spectral indices between $-1.7$
1234: and $-1.8$ in a study of Ulysses magnetic-field data covering a range
1235: of heliographic latitudes and radii.
1236:
1237: Overall, the spectra are more consistent with a Kolmogorov scaling
1238: than an Iroshnikov-Kraichnan scaling. It should be emphasized,
1239: however, that the observations in several cases are consistent with
1240: inertial range spectra that are steeper than a Kolmogorov spectrum.
1241: Spectral indices~$> 5/3$ have been found in previous theoretical
1242: studies of weak incompressible MHD turbulence (Bhattacharjee \& Ng
1243: 1997, Goldreich \& Sridhar~1997, Galtier et~al~2000, Bhattacharjee \&
1244: Ng~2001, Perez \& Boldyrev~2008), as well as strong isotropic MHD
1245: turbulence with cross helicity (Grappin et~al~1983). In this paper, it
1246: is argued that spectral indices~$> 5/3$ are a consequence of cross
1247: helicity in strong anisotropic MHD turbulence.
1248:
1249: The simplest way to apply this paper to solar wind turbulence is to
1250: model the solar wind fluctuations at some location as steady-state,
1251: forced, homogeneous turbulence with the same average value of
1252: $w_{k_f}^+/w_{k_f}^-$, where $w_{k_f}^+$ and $w_{k_f}^-$ are the rms
1253: amplitudes of the $w^\pm$ fluctuations at the outer scale~$k_f^{-1}$.
1254: Upon setting $w_{k_\perp}^\pm \propto k_\perp ^{-a^\pm}$, one can
1255: write $(w^+_{k_\perp}/w^-_{k_\perp})^2 \propto (k_\perp/k_d)^{-2a^+ +
1256: 2a^-}$, where it is assumed that the spectra are equal at the
1257: dissipation wave number~$k_d$. Equation (\ref{eq:apam1}) then gives
1258: $(w^+_{k_\perp}/w^-_{k_\perp})^2 \simeq (k_\perp/k_d)^{1-3a^+}$, and
1259: the value of $a^+$ can be obtained from the equation
1260: $(w^+_{k_f}/w^-_{k_f})^2 \simeq (k_f/k_d)^{1-3a^+}$. The total-energy
1261: spectrum, $E(k_\perp) = E^+(k_\perp) + E^-(k_\perp)$, is
1262: approximately~$k_\perp^{-1} (w_{k_\perp}^+)^2$, although it is flatter
1263: than $k_\perp^{-1} (w_{k_\perp}^+)^2$ near the dissipation scale where
1264: the flatter spectrum of the~$w^-$ fluctuations is important. Thus, at
1265: scales much larger than the dissipation scale,
1266: \begin{equation}
1267: E(k_\perp) \propto k_\perp^{-q},
1268: \end{equation}
1269: where
1270: \begin{equation}
1271: q = \frac{5}{3} + \frac{2 \log_{10}[(w_{k_f}^+)^2/ (w_{k_f}^-)^2]}{3\log_{10}(k_d/k_f)}.
1272: \label{eq:defc3}
1273: \end{equation}
1274: Upon defining
1275: the outer-scale fractional cross helicity as
1276: \begin{equation}
1277: \sigma_c = \frac{(w_{k_f}^+)^2 - (w_{k_f}^-)^2}{(w_{k_f}^+)^2 + (w_{k_f}^-)^2}
1278: \label{eq:defsigmac}
1279: \end{equation}
1280: one can rewrite equation~(\ref{eq:defc3}) as
1281: \begin{equation}
1282: q = \frac{5}{3} + \frac{2 \log_{10}[(1+ \sigma_c)/ (1- \sigma_c)]}{3\log_{10}(k_d/k_f)}.
1283: \label{eq:defc32}
1284: \end{equation}
1285: The spectral index from equation~(\ref{eq:defc32}) is plotted in
1286: Figure~\ref{fig:f5}, assuming that $k_d/k_f = f_d/f_b = 3780$, where
1287: $f_b = (3.5 \mbox{ hours})^{-1}$ is the break frequency at 1~AU
1288: discussed above (Matthaeus \& Goldstein 1986) and $f_d = 0.3 \mbox{
1289: s}^{-1}$ is the frequency at the dissipation scale. (Smith
1290: et~al~2006) With this choice, $q=1.78$ for $(w_{k_f}^+)^2/ (w_{k_f}^-)^2
1291: = 4$ and $q=1.74$ for $(w_{k_f}^+)^2/ (w_{k_f}^-)^2 = 2$. When
1292: equation~(\ref{eq:defc32}) is applied to the solar wind, $\sigma_c$
1293: should be interpreted as the cross helicity at the outer
1294: scale~$k_f^{-1}$ averaged over at least a few outer-scale
1295: fluctuations.
1296:
1297:
1298: \begin{figure}[h]
1299: \includegraphics[width=3.5in]{f5.eps}
1300: \caption{\footnotesize Dependence of spectral index of the total
1301: energy spectrum,~$q$, on the outer-scale fractional cross
1302: helicity~$\sigma_c$. The ratio of the dissipation wavenumber~$k_d$
1303: to the perpendicular wavenumber at the outer-scale~$k_f$ in
1304: equation~(\ref{eq:defc32}) is taken to be~$3780$.
1305: \label{fig:f5}}
1306: \end{figure}
1307:
1308: Some caution, however, is warranted when applying
1309: equation~(\ref{eq:defc32}) to the solar wind because solar wind
1310: conditions vary with~$r$, while equation~(\ref{eq:defc32}) is based on
1311: results for homogeneous turbulence. The condition under which it is
1312: valid to treat the solar-wind fluctuations at a some~$r$ with an
1313: observed~$\sigma_c$ as homogeneous turbulence with the same $\sigma_c$
1314: is that the cascade time at the outer scale $\tau_{k_f}^\pm$ be much
1315: smaller than the time $t_{\rm adv} = r/U$ for turbulent structures to
1316: be advected a distance~$r$, where $U$ is the solar wind speed.
1317: However, this condition is often either violated or only marginally
1318: satisfied. This is illustrated by the results of Matthaeus \&
1319: Goldstein (1982) based on four days of Voyager data (in their
1320: Table~1). For the fluctuations at 1~AU, these authors found an rms
1321: velocity fluctuation of~$\delta v = 25.5$~km/s [which is comparable
1322: to~$v_A$ at 1~AU], a velocity correlation length (outer scale) of
1323: $L_\perp \sim 2.83\times 10^{11}$~cm, and an average solar-wind speed
1324: of $352$~km/s. In this particular data set, $\sigma_c = 0.06$, so
1325: that $w_{k_f}^+ \sim w_{k_f}^-$, with $\tau_{k_f}^+ \sim \tau_{k_f}^-
1326: \sim L_\perp /\delta v = 1.11 \times 10^5$~s. On the other hand,
1327: $t_{\rm adv} = \mbox{ (1 AU)}/U = 4.26 \times 10^5$~s, which
1328: marginally satisfies the condition $t_{\rm adv} \gg \tau_{k_f}^+$.
1329: However, if $w_{k_f}^+ /w_{k_f}^-$ were instead~$\sim 3$ at 1~AU, as
1330: in the results of Bavassano et~al~(2000), then $\tau_{k_f}^+$ would
1331: be somewhat larger than~$t_{\rm adv}$. Thus, the inhomogeneity of
1332: the solar wind may influence the effects of cross helicity on the
1333: spectral indices. However, more detailed modeling of inhomogeneous
1334: solar-wind turbulence is beyond the scope of this paper.
1335:
1336: \section{Comparison to Other Studies}
1337: \label{sec:comp}
1338:
1339: In this section, the results of this paper are compared to
1340: two recent studies of strong anisotropic incompressible MHD turbulence
1341: with cross helicity.
1342:
1343: \subsection{Lithwick, Goldreich, \& Sridhar (2007) }
1344:
1345: The model for the cascade of energy to larger~$k_\parallel$ used in
1346: this paper is based on the results of Lithwick, Goldreich, \&
1347: Sridhar~(2007) (hereafter LGS07). As a result, in both studies, if
1348: $w^+$ and $w^-$ have comparable correlation lengths in the direction
1349: of~$\bm{B}_0$ at the outer scale, then $\overline{
1350: k}_\parallel^{\,+}\simeq \overline{ k}_\parallel^{\,-}$ at all
1351: smaller scales. The principal difference between this paper and~LGS07
1352: lies in our treatment of the cascade time for the dominant fluctuation
1353: type,~$w^+$. LGS07 argue that if $w_{k_\perp}^+ \gg w_{k_\perp}^-$,
1354: $\chi_{k_\perp}^+ \sim 1$, and $\chi_{k_\perp}^- \ll 1$, then the
1355: shearing applied by $w^-$ wave packets on a $w^+$ wave packet at
1356: perpendicular scale~$k_\perp^{-1}$ is coherent over a time~$(k_\perp
1357: w_{k_\perp}^-)^{-1}$, which greatly exceeds the time $ (\overline{
1358: k}_\parallel^{\,-} v_A)^{-1}$ required for a $w^+$ and $w^-$ wave
1359: packet at perpendicular scale $k_\perp^{-1}$ to pass through each
1360: other. In contrast, in this paper, it is argued that the coherence
1361: time for the straining of the $w^+$ wave packet is of order the
1362: ``crossing time'' $ (\overline{ k}_\parallel^{\,-} v_A)^{-1}$. As a
1363: result, the results obtained in this paper for the inertial-range
1364: power spectra, degree of anisotropy, cascade time, and energy fluxes
1365: are different from those of~LGS07.
1366:
1367: The approach taken in this paper is motivated by the following
1368: argument. As argued by LGS07 and Maron \& Goldreich~(2001), the $w^-$
1369: fluctuations propagate approximately along the hypothetical magnetic
1370: field lines obtained from the sum of~$\bm{B}_0$ and the magnetic field
1371: of the~$w^+$ fluctuations. Let us call these hypothetical magnetic
1372: field lines the ``$w^+$ field lines,'' and let us consider a $w^+$
1373: wave packet of perpendicular scale~$k_\perp^{-1}$ and parallel scale
1374: $(\overline{ k}_\parallel)^{-1}$, where $\overline{ k}_\parallel =
1375: \overline{ k}_\parallel^+ \simeq \overline{ k}_\parallel^-$. Let us
1376: work in a frame of reference that moves at speed~$v_A$ in the~$-z$
1377: direction along with the~$w^+$ fluctuations. Let us also take an
1378: initial snapshot of the turbulence at $t=0$ and trace out all of the
1379: ``$w^+$ field lines'' that pass through our wave packet. The volume
1380: filled by these $w^+$~field lines is the ``source region'' from which
1381: the $w^-$ wavepackets encountered by our $w^+$ wave packet
1382: originate. If we wait one crossing time $(\overline{ k}_\parallel
1383: v_A)^{-1}$ and take a new snapshot of the turbulence, then at any
1384: given location the $w^+$ fluctuations will not have changed very much,
1385: since $w_{k_\perp}^- \ll w_{k_\perp}^+$. However, if we trace out the
1386: new $w^+$ field lines passing through our wave packet, the volume that
1387: is filled by these new $w^+$ field lines will differ substantially
1388: from the initial source region at distances~$\ga (\overline{
1389: k}_\parallel)^{-1} $ from our wave packet due to the rapid
1390: divergence of neighboring field lines in MHD turbulence. In other
1391: words, small local changes in $w^+$ lead to large changes in the
1392: connectivity of the $w^+$ field lines.
1393:
1394: To see this, let the $w^+$ field line that passes through some
1395: point~$P$ in our wave packet at~$t=0$ be called ``field line~$A$.''
1396: Let the $w^+$ field line that passes through point~$P$ at $t=
1397: (\overline{ k}_\parallel v_A)^{-1}$ be called ``field line~$B$.'' As
1398: before, let us work in a frame of reference that moves at speed~$v_A$
1399: in the~$-z$ direction. Field lines~$A$ and~$B$ are fixed curves, since
1400: they are traced out within two snapshots of the turbulence. If we
1401: follow field line~$B$ for a distance~$\ll (\overline{
1402: k}_\parallel)^{-1}$, it will separate from field line~$A$ by some
1403: small distance~$x$ that is~$\ll k_\perp^{-1}$. If we continue to
1404: follow field line~$B$, its separation from field line~$A$ is analogous
1405: to the separation of two neighboring field lines within a single
1406: snapshot of the turbulence. As shown by Narayan \& Medvedev~(2001),
1407: Chandran \& Maron~(2004), and Maron, Chandran, \& Blackman~(2004), if
1408: a pair of field lines is separated by a distance~$x \ll k_\perp^{-1}$
1409: at one location, then the distance the field-line pair must be
1410: followed before it separates by a distance~$k_\perp^{-1}$ is a few
1411: times the parallel size of an eddy of perpendicular
1412: size~$k_\perp^{-1}$ - i.e., a few times~$(\overline{
1413: k}_\parallel)^{-1}$. It turns out that the particular value of $x$
1414: has little effect unless one considers the (irrelevant) case in which
1415: $x/d \sim e^{-N},$ where $N$ is large and $d$ is the perpendicular
1416: dissipation scale. (Chandran \& Maron 2004) This is because within the
1417: inertial range the amount of magnetic shear increases towards small
1418: scales; therefore, if $x$ is made very small, then the distance one
1419: has to follow the field-line pair in order for~$x$ to double becomes
1420: very small. Thus, as a result of the rapid divergence of neighboring
1421: field lines in MHD turbulence, the ``source region'' of our $w^+$ wave
1422: packet at $t=(\overline{ k}_\parallel v_A)^{-1}$ differs substantially
1423: from the source region at $t=0$ at distances $\ga (\overline{
1424: k}_\parallel)^{-1}$ from our wave packet. Because the $w^-$
1425: fluctuations vary rapidly in the direction perpendicular to the
1426: magnetic field, the $w^-$ wave packets encountered by our $w^+$ wave
1427: packet will decorrelate on a time scale of order the crossing time
1428: $(\overline{ k}_\parallel v_A)^{-1}$ due to the time evolution of the
1429: source region.
1430:
1431: It should noted that there are two unexplained aspects of LGS07's
1432: model, as pointed out by Beresnyak \& Lazarian~(2007). The first
1433: concerns the nature of the transition from the weak-turbulence
1434: regime~($\chi_{k_\perp}^+ \ll 1$ and $\chi_{k_\perp}^- \ll 1$) to the
1435: strong-turbulence regime~($\chi_{k_\perp}^+\sim 1$ and
1436: $\chi_{k_\perp}^- \ll 1$). In LGS07's analysis, as one passes from
1437: the weak regime to the strong regime, the coherence time for the
1438: straining of the $w^+$ wave packets by~$w^-$ wave packets increases by
1439: a factor of $(\chi_{k_\perp}^-)^{-1}$ and the energy cascade
1440: time~$\tau_{k_\perp}^+$ decreases by a factor of~$\chi_{k_\perp}^-$.
1441: It is not clear why these large changes should occur across the
1442: transition scale. The second issue is that LGS07 find that
1443: $E^+(k_\perp) \propto k_\perp^{-5/3}$ and $E^-(k_\perp) \propto
1444: k_\perp^{-5/3}$ regardless of the fractional cross helicity. Since
1445: the ratio $E^+(k_\perp)/E^-(k_\perp)$ is independent of wavenumber, it
1446: is not clear how pinning could occur in their model, or how the
1447: spectra would behave near the dissipation scale.
1448:
1449:
1450: \subsection{Beresnyak \& Lazarian (2007)}
1451:
1452: Beresnyak \& Lazarian (2007) (hereafter BL07) have published an online
1453: article on strong MHD turbulence with cross helicity. The following
1454: discussion refers to the version of their article that is available
1455: electronically as of the writing of this paper. The work of BL07 is
1456: similar to this paper in that both studies take the dominant
1457: fluctuation type, $w^+$, to undergo a weak cascade. Also,
1458: equations~(\ref{eq:defapam}) through (\ref{eq:olkp1}) of this paper
1459: are equivalent to their equation~(5), except for the fact that they
1460: take the parallel correlation lengths of $w^+$ and $w^-$ to differ by
1461: a constant multiplicative factor when a power-law solution for the
1462: spectra is assumed. (The possibility of more general solutions is
1463: claimed by BL07.) On the other hand, there are a number of significant
1464: differences between this paper and Beresnyak \& Lazarian's (2007)
1465: work. They argue that for the $w^+$ fluctuations, the dominant
1466: nonlinear interactions are between fluctuations with comparable
1467: parallel correlation lengths and different perpendicular correlation
1468: lengths, whereas for the $w^-$ fluctuations the dominant interactions
1469: are between fluctuations with comparable perpendicular scales.
1470: Here, it is argued that for both $w^+$ and $w^-$ the
1471: dominant interactions are between fluctuations with similar
1472: perpendicular scales. When $w^+_{k_\perp} \gg w^-
1473: _{k_\perp}$ and $\chi_{k_\perp}^+ \sim 1$, their Figure~1 suggests
1474: that the parallel correlation length of~$w^+$ fluctuations can be less
1475: than the parallel correlation length of~$w^-$ fluctuations. It is
1476: argued in section~\ref{sec:unequal} of this paper that this can not be
1477: the case, because the $w^+$ fluctuations will imprint their parallel
1478: correlation length onto the~$w^-$ fluctuations. They argue that the
1479: scalings given by equations~(\ref{eq:defapam}) and (\ref{eq:olkp1})
1480: [equivalently, their equation~(5)] can not apply if the parallel
1481: correlation lengths of the $w^+$ and $w^-$ fluctuations are equal,
1482: arguing that this would require $\epsilon^+ = \epsilon^-$, whereas in
1483: this paper the ratio~$\epsilon^+/\epsilon^-$ depends upon the slopes
1484: of the power spectra, as in weak turbulence. They argue that if the
1485: $w^+$ and $w^-$ fluctuations are driven with the same parallel
1486: correlation length at the outer scale, there will be a non-power law
1487: part of the solution at large scales that will transition at smaller
1488: scales to a power law solution with $w_{k_\perp}^+$ and
1489: $w^-_{k_\perp}$ both $\propto k_\perp^{-1/3}$ and $\overline{
1490: k}_\parallel ^{\,\pm}\propto k_\perp^{2/3}$. In contrast, in this
1491: paper a power-law solution starting at the outer scale is obtained
1492: with different scalings for $w^+_{k_\perp}$ and $w^-_{k_\perp}$, and
1493: with $\overline{ k}_\parallel^{\pm}$ growing more slowly
1494: than~$k_\perp^{2/3}$.
1495:
1496: \section{Conclusion}
1497:
1498: This paper proposes a new phenomenology for strong, anisotropic,
1499: incompressible MHD turbulence with cross helicity and introduces a
1500: nonlinear advection-diffusion equation [equation~(\ref{eq:FPpm})] to
1501: describe the time evolution of the anisotropic power spectra of the
1502: $w^+$ and $w^-$ fluctuations. It is found that in steady state the
1503: one-dimensional power spectra of the energetically dominant $w^+$
1504: fluctuations, $E^+(k_\perp)$, is steeper than $k_\perp^{-5/3}$, and
1505: that $E^+(k_\perp)$ becomes increasingly steep as the fractional cross
1506: helicity~$\sigma_c$ increases. Increasing $\sigma_c$ also increases
1507: the energy cascade time of the $w^+$ fluctuations, reduces the
1508: turbulent heating power for a fixed fluctuation energy, and increases
1509: the anisotropy of the fluctuations at small scales.
1510:
1511: Although most of the discussion has focused on forced, steady-state
1512: turbulence, the results of this paper can also be applied to decaying
1513: turbulence. For example, equations~(\ref{eq:tau1}) and
1514: (\ref{eq:tau2}) can be used to estimate the time scale for turbulence
1515: to decay. The resulting prediction is that if the fluctuations are
1516: initially excited with $w_{k_f}^+ \gg w_{k_f}^-$ and with comparable
1517: parallel correlation lengths at the outer scale, then the turbulence
1518: will decay into a state in which~$w^-$ fluctuations are absent, as in
1519: the earlier work of Dobrowolny, Mangeney, \& Veltri (1980), Grappin
1520: et~al~(1983), and Lithwick \& Goldreich (2003). This ``maximally
1521: aligned'' state will then be free from nonlinear interactions, and
1522: will persist for long times until it damps via linear dissipation.
1523:
1524:
1525: \acknowledgements I thank Chuck Smith, Sebastien Galtier, Bernie
1526: Vasquez, Yoram Lithwick, Alex Lazarian, and Andrey Beresnyak for
1527: helpful comments and suggestions. This work was supported in part by
1528: the NSF/DOE Partnership in Basic Plasma Science and Engineering under
1529: grant number AST-0613622, by NASA under grant numbers NNX07AP65G and
1530: NNH06ZDA001N-SHP06-0071, and by DOE under grant number
1531: DE-FG02-07-ER46372.
1532:
1533:
1534:
1535:
1536: \references
1537:
1538:
1539: Antonucci, E., Dodero, M. A., \& Giordano, S. 2000, Sol. Phys.,
1540: 197, 115
1541:
1542: Barnes, A. 1966, Phys. Fluids, 9, 1483
1543:
1544: Bavassano, B., Pietropaolo, E., \& Bruno, R. 2000, J. Geophys. Res.,
1545: 105, 15959
1546:
1547: Beresnyak, A., \& Lazarian, A. 2006, ApJ, 640, L175
1548:
1549: Beresnyak, A., \& Lazarian, A. 2007, arXiv:0709.0554v1
1550:
1551: Bhattacharjee, A. \& Ng, C. S. 2001, ApJ 548, 318
1552:
1553: Bieber, J., Matthaeus, W., Smith, C., Wanner, W., Kallenrode, M., \&
1554: Wibberenz, G. 1994, ApJ, 420, 294
1555:
1556: Biskamp, D., Schwarz, E., \& Drake, J. F. 1996, Phys. Rev. Lett., 76, 1264
1557:
1558: Biskamp, D., Schwarz, E., Zeiler, A., Celani, A., \& Drake, J. F. 1999, Phys. Plasmas, 6, 751
1559:
1560: Boldyrev, S., Nordlund, A., \& Padoan, P. 2002, Phys. Rev. Lett., 89,
1561: 031102
1562:
1563: Boldyrev, S. 2005, ApJ, 626, L37
1564:
1565: Boldyrev, S. 2006, Phys. Rev. Lett., 96 115002
1566:
1567: Brodin, G., Stenflo, L., \& Shukla, P. K. 2006, Solar Phys., 236,
1568: 285
1569:
1570: Bruno, R., \& Carbone, V. 2005, Living Rev. Solar Phys., 2, 4
1571: [Online article: cited Nov. 14, 2007, http://www.livingreviews.org/lrsp-2005-4]
1572:
1573: Chandran, B. 2004, Sp. Sci. Rev., 292, 17
1574:
1575: Chandran, B. 2005, Phys. Rev. Lett., 95, 265004
1576:
1577: Cho, J., \& Vishniac, E. 2000, ApJ, 539, 273
1578:
1579: Cho, J., Lazarian, A., \& Vishniac, E. 2002, ApJ, 564, 291
1580:
1581: Cho, J., \& Lazarian, A. 2002, Phys. Rev. Lett., 88, 245001
1582:
1583: Cho, J., \& Lazarian, A. 2003, MNRAS, 345, 325
1584:
1585: Cho, J., \& Lazarian, A. 2004, ApJL, 615, L41
1586:
1587: Cranmer, S. R. \& van Ballegooijen, A. A. 2005, ApJS, 156, 265
1588:
1589: Cranmer, S. R. \& van Ballegooijen, A. A. 2007, ApJS, 171, 520
1590:
1591: Dmitruk, P., Milano, L. J., \&
1592: Matthaeus, W. H. 2001, ApJ, 548, 482
1593:
1594: Dmitruk, P., Matthaeus, W. H., Milano, L. J., Oughton, S., Zank, G. P.,
1595: \& Mullan, D. J. 2002, ApJ, 575, 571
1596:
1597: Dobrowolny, M., Mangeney, A., Veltri, P. L. 1980, Phys. Rev. Lett., 35, 144
1598:
1599: Elmegreen, B. G., \& Scalo, J., ARAA, 42, 211
1600:
1601: Galtier, S., Nazarenko, S. V., Newell, A. C., \& Pouquet,
1602: A. 2000, J. Plasma Phys., 63, 447
1603:
1604: Galtier, S., \& Chandran, B. 2006, Phys. Plasmas, 3, 114505
1605:
1606: Galtier, S., \& Buchlin, E. 2007, ApJ, 656, 560
1607:
1608: Goldreich, P., \& Sridhar, S. 1995, ApJ, 438, 763
1609:
1610: Goldreich, P., \& Sridhar, S. 1997, ApJ, 485, 680
1611:
1612: Goldstein, B. E., Smith, E. J., Balogh, A., Horbury, T. S., Goldstein, M. L.,
1613: \& Roberts, D. A. 1995, Geophys. Res. Lett., 22, 3393
1614:
1615: Grappin, R., Pouquet, A., \& L\'eorat, J. 1983, A\&A, 126, 51
1616:
1617: Haugen, N. E. L., Brandenburg, A., \& Dobler, W., Phys. Rev. E. 2004,
1618: 70, 016308
1619:
1620: Higdon, J. 1984, ApJ, 285, 109
1621:
1622: Howes, G. G., Cowley, S. C., Dorland, W., Hammett, G. W., Quataert, E.,
1623: \& Schekochihin, A. A. 2006, ApJ, 651 590-614
1624:
1625: Howes, G. G., Cowley, S. C., Dorland, W., Hammett, G. W., Quataert, E.,
1626: \& Schekochihin, A. A. 2007a, arXiv:0707.3147
1627:
1628: Howes, G. G., Dorland, W., Cowley, S. C., Hammett, G. W., Quataert, E.,
1629: Schekochihin, A. A., \& Tatsuno, T. 2007b, arXiv:0711.4355
1630:
1631: Iroshnikov, P. 1963, Astron. Zh. 40, 742
1632:
1633: Klein, L. W., Matthaeus, W. H., Roberts, D. A., \& Goldstein, M. L. 1992,
1634: {\em Solar Wind Seven; Proceedings of the 3rd COSPAR Colloquium, Goslar, Germany,} pp. 197-200.
1635:
1636: Kohl, J. L. et al 1998, ApJL, 501, 127
1637:
1638: Kolmogorov, A. N. 1941, Dokl. Akad. Nauk SSSR, 30, 299
1639:
1640: Kraichnan, R. H. 1965, Phys. Fluids 8, 1385
1641:
1642: Kuznetsov, E. A. 2001, J. Exp. Theor. Phys., 93, 1052
1643:
1644: Leith, C. E., \& Kraichnan, R. H. 1972, J. Atmos. Sci., 29, 1041
1645:
1646: Lithwick, Y., \& Goldreich, P. 2001, ApJ, 562, 279
1647:
1648: Lithwick, Y., \& Goldreich, P. 2003, ApJ, 582, 1220
1649:
1650: Lithwick, Y., Goldreich, P., \& Sridhar, S. 2007, ApJ, 655, 269 (LGS07)
1651:
1652: Luo, Q., \& Melrose, D. 2006, MNRAS, 368, 1151
1653:
1654: Maron, J., Chandran, B., \& Blackman, E. 2004, Phys. Rev. Lett., 92,
1655: 045001
1656:
1657: Maron, J., \& Goldreich, P. 2001, ApJ, 554, 1175
1658:
1659: Marsch, E., \& Tu C.-Y. 1996, J. Geophys. Res., 101, 11149
1660:
1661: Mason, J., Cattaneo, F., \& Boldyrev, S. 2006, Phys. Rev. Lett., 97, 255002
1662:
1663: Matthaeus, W. H., \& Goldstein, M. L. 1982, J. Geophys. Res., 87, 6011
1664:
1665: Matthaeus, W. H., \& Goldstein, M. L. 1986, Phys. Rev. Lett., 57, 495
1666:
1667: Matthaeus, W. H., \& Montgomery, D. 1980, New York Acad. Sci., 357, 203
1668:
1669: Matthaeus, W. H., Zank, G. P., Leamon, R. J.,
1670: Smith, C. W., Mullan D. J., \& Oughton, S. 1999, Sp. Sci. Rev., 87, 269
1671:
1672: Matthaeus, W. H., Mullan D. J., Dmitruk, P., Milano, L., \& Oughton, S. 2002,
1673: Non. Proc. Geophys., 9, 1
1674:
1675: Matthaeus, W. H., Dmitruk, P., Smith, D., Ghosh, S., \& Oughton, S. 2003,
1676: Geophys. Res. Lett., 30, 4-1
1677:
1678: Matthaeus, W. H., Pouquet, A., Mininni, P. D., Dmitruk, P.,
1679: \& Breech, B. 2007, arXiv:0708.0801
1680:
1681: Mininni, P., \& Pouquet, A. 2007, Phys. Rev. Lett., 99, 254502
1682:
1683: Moffatt, H. K. 1978, {\em Magnetic field generation in electrically conducting fluids} (Cambridge, England: Cambridge University Press, 1978)
1684:
1685: Montgomery, D., \& Matthaeus, W. 1995, ApJ, 447, 706
1686:
1687: Montgomery, D., \& Turner, L. 1981, Phys. Fluids, 24, 825
1688:
1689: M\"uller, W. C., \& Biskamp, D. 2000, Phys. Rev. Lett., 84, 475
1690:
1691: M\"uller, W. C., Biskamp, D., \& Grappin, R. 2003, Phys. Rev. E, 67, 066302
1692:
1693: M\"uller, W. C., \& Grappin, R. 2005, Phys. Rev. Lett., 95, 114502
1694:
1695: Ng, C. S., \& Bhattacharjee, A. 1996, ApJ, 465, 845
1696:
1697: Ng, C. S., \& Bhattacharjee, A. 1997, Phys. Plasmas, 4, 605
1698:
1699: Oughton, S., Matthaeus, W. H., \& Ghosh, S. 1995, LNP, Vol.~462: {\em
1700: Small-Scale Structures in Three-Dimensional Hydrodynamic and
1701: Magnetohydrodynamic Turbulence}, p. 273
1702:
1703: Oughton, S., Dmitruk, P., \& Matthaeus, W. H. 2006, Phys. Plasmas,
1704: 13, 2306
1705:
1706: Padoan, P., Jimenez, R., Nordlund, A., \& Boldyrev, S. 2004, Phys.
1707: Rev. Lett., 92, 191102
1708:
1709: Perez, J. C., \& Boldyrev, S. 2008, ApJL, 672, L61
1710:
1711: Podesta, J. J., Roberts, D. A., \& Goldstein, M. L. 2007, ApJ, 664, 543
1712:
1713: Pouquet, A., Sulem, P. L., \& Meneguzzi, M. 1988, Phys. Fluids, 31, 2635
1714:
1715: Roberts, D. A., Goldstein, M. L., Klein, L. W., \& Matthaeus, W. H. 1987,
1716: J. Geophys. Res., 92, 12023
1717:
1718: Schekochihin, A. A., Cowley, S. C., Dorland, W., Hammett, G. W., Howes, G. G., Quataert, E., \& Tatsuno, T 2007b, arXiv:0704.0044
1719:
1720: Shebalin, J. , Matthaeus, W., \& Montgomery, D. 1983, J. Plasma Phys.,
1721: 29, 525
1722:
1723: Shukla, P. K., Brodin, G., \& Stenflo, L. 2006, Phys. Lett. A,
1724: 353, 73
1725:
1726: Skilling, J., McIvor, I., \& Holmes, J. 1974, MNRAS, 167, 87P
1727:
1728: Smith, C. W. 2003, {\em Solar Wind Ten: Proceedings of the Tenth
1729: International Solar Wind Conference}, AIP Conference Proceedings,
1730: 679, 413
1731:
1732: Smith, C. W., Hamilton, K., Vasquez, B., \& Leamon, R. 2006, ApJL, 645, 85
1733:
1734: Stone, J. M., Ostriker, E. C., \& Gammie, C. F. 1998, ApJL, 508, 99
1735:
1736: Taylor, G. I. 1938, Proc. R. Soc. London A, 164, 476
1737:
1738: Velli, M. 1993, A\&A, 270, 304
1739:
1740: Verdini, A., \& Velli, M. 2007, ApJ, 662, 669
1741:
1742: Zhou, Y., \& Matthaeus, W. H. 1990, J. Geophys. Res., 95, 10291
1743:
1744: \end{document}
1745:
1746:
1747:
1748: