1: \documentclass{emulateapj}
2: %\documentclass[12pt,preprint]{aastex}
3: %\usepackage{amsmath}
4:
5: \newcommand{\Msun}{\ensuremath{M_\odot}}
6: %\newcommand{\boldsymbol}{\boldmath}
7: \newcommand{\boldsymbol}[1]{\mbox{\boldmath{$#1$}}}
8:
9: \shorttitle{Evolution of Magnetic Turbulence in Shocks}
10: \shortauthors{Chang, Spitkovsky, \& Arons}
11:
12: \begin{document}
13:
14: \title{Long Term Evolution of Magnetic Turbulence in Relativistic Collisionless Shocks: Electron-Positron Plasmas}
15:
16: \author{Philip Chang\altaffilmark{1,2,3}, Anatoly Spitkovsky\altaffilmark{4}, and Jonathan Arons\altaffilmark{1,2,5,6}}
17:
18: \altaffiltext{1} {Department of Astronomy, Campbell Hall, University
19: of California, Berkeley, CA 94720; pchang@astro.berkeley.edu,
20: arons@astro.berkeley.edu}
21: \altaffiltext{2} {Theoretical Astrophysics Center}
22: \altaffiltext{3} {Miller Institute for Basic Research}
23: \altaffiltext{4} {Department of Astrophysical Sciences, Peyton Hall,
24: Princeton University, Princeton, NJ: anatoly@astro.princeton.edu}
25: \altaffiltext{5} {Department of Physics, LeConte Hall, University
26: of California, Berkeley, CA 94720}
27: \altaffiltext{6}{Kavli Institute for Particle Astrophysics and Cosmology, Stanford
28: University}
29:
30: \begin{abstract}
31: We study the long term evolution of magnetic fields generated by a
32: collisionless relativistic $e^+e^-$ shock which is initially
33: unmagnetized. Our 2D particle-in-cell numerical simulations show
34: that downstream of such a Weibel-mediated shock, particle
35: distributions are close to isotropic, relativistic Maxwellians, and
36: the magnetic turbulence is highly intermittent spatially. The
37: non-propagating magnetic fields in the turbulence form relatively isolated regions
38: with transverse dimension $\sim 10-20$ skin depths. These structures
39: decay in amplitude, with little sign of downstream merging. The
40: fields start with magnetic energy density $\sim (0.1-0.2)$ of the
41: upstream kinetic energy within the shock transition, but rapid
42: downstream decay drives the fields to much smaller values, below
43: $10^{-3}$ of equipartition after $\sim 10^3$ skin depths.
44:
45: % At late times in the simulations, which last for
46: % $\sim 4000 \omega_p^{-1}$ and follow the motions of $15 \time 10^9$
47: % particles, where $\omega_p$ is the relativistic plasma frequency
48: % based on upstream parameters, the total magnetic energy, averaged
49: % over the transverse dimension, declines in proportion to $t^{-1}$.
50: % Noise due to the finite number of particles limits following this
51: % decay to magnetic energy densities greater than $0.001 nkT$.
52:
53: In an attempt to construct a theory that follows field decay to
54: these smaller values, we explore the hypothesis that the observed
55: damping is a variant of Landau damping in an unmagnetized plasma.
56: The model is based on the small value of the downstream magnetic
57: energy density, which
58: suggests that particle orbits are only weakly perturbed from
59: straight line motion, if the turbulence is homogeneous. Using
60: linear kinetic theory applied to electromagnetic fields in an
61: isotropic, relativistic Maxwellian plasma, we find a simple analytic
62: form for the damping rates, $\gamma_k$, in two and three dimensions
63: for small amplitude, subluminous electromagnetic fields. We find
64: that magnetic energy does damp due to phase mixing of current
65: carrying particles as $(\omega_p t)^{-q}$ with $q \sim 1$. This
66: overall decay compares well to that found in simulations, since it
67: depends primarily on the longest wavelength modes, $kc/\omega_p \ll
68: 1$. However, the theory predicts overly rapid damping of short
69: wavelength modes. We speculate that magnetic trapping of a
70: substantial fraction of the particles within the highly spatially
71: intermittent downstream magnetic structures may be the origin of
72: this discrepancy. In addition, trapping may form the basis for
73: MHD-like behavior, permitting a small fraction of the initial
74: magnetic energy to persist for times much greater than have been
75: followed in the simulations.
76:
77: We briefly speculate on other physical processes, which depend on
78: the presence of suprathermal particles, that may cause the
79: generation of longer wavelength magnetic fields that create a
80: magnetized plasma ($kr_{Larmor} \ll 1$), in which the damping is not
81: as fast. However, absent such additional physical processes, we
82: conclude that initially unmagnetized relativistic shocks in
83: electron-positron plasmas are unable to form persistent downstream
84: magnetic fields. These results put interesting constraints on
85: synchrotron models for the prompt and afterglow emission from GRBs.
86: We also comment on the relevance of these results for relativistic
87: shocks in electron-ion plasmas.
88:
89: % Because
90: % the downstream magnetic turbulence actually is highly intermittent
91: % in space, magnetic trapping of a substantial fraction of the
92: % particles within the magnetic structures occurs and may be the
93: % origin of the discrepancy between the theory and the late time
94: % damping observed numerically. We speculate that trapping may allow
95: % persistence of a small fraction of the initial magnetic energy to
96: % times much greater than have been followed in the simulations, since
97: % permanently magnetically trapped particles are the basis for MHD
98: % behavior.
99:
100:
101: % Because the downstream magnetic turbulence actually is highly
102: % intermittent in space, magnetic trapping of a substantial fraction
103: % of the particles within the magnetic structures occurs and may be
104: % the origin of the discrepancy between the theory and the late time
105: % damping observed numerically. We exhibit results from the
106: % simulations which demonstrate that magnetic trapping does occur in
107: % the downstream medium, since even though the average magnetic energy
108: % density is small compared to the thermal energy density, the actual
109: % fields in the widely space filaments are strong enough to deflect
110: % and trap a large fraction of the particles that encounter the
111: % filaments. We speculate that trapping may allow persistence of a
112: % small fraction of the initial magnetic energy to times much greater
113: % than have been followed in the simulations, since permanently
114: % magnetically trapped particles are the basis for MHD behavior.
115:
116: \end{abstract}
117:
118: \keywords{shock waves -- turbulence -- gamma ray: bursts -- plasmas}
119:
120: \section{Introduction}
121:
122: The prompt emission and afterglows of gamma-ray bursts (GRBs) may be
123: manifestations of ultrarelativistic shock waves, propagating in media
124: where the large scale upstream magnetization is too weak to affect the
125: shock structure, and too weak, if simply compressed by the shock, to
126: provide the magnetization inferred from synchrotron models of the
127: burst emission (see Piran 2005ab and references therein). The weakly
128: magnetized outflows in the rotational equators of rotation powered
129: pulsars (Coroniti 1990) may also be sites of essentially unmagnetized
130: shock waves terminating the relativistic winds in a region which
131: occupies a finite latitude band with respect to the rotational equator
132: of the underlying neutron star. Yet the radiation from these systems
133: has been modeled as being due to synchrotron radiation, which requires
134: the presence of a magnetic field strong enough to deflect particles
135: through a substantial fraction of their Larmor orbits. At the very
136: least, this requires magnetic structures with amplitude, $\delta B$,
137: whose characteristic dimensions, $R_B$, are comparable to or larger
138: than $mc^2 /e \delta B$, for all
139: particle energies inferred to contribute to the observed radiation.
140:
141: Unmagnetized anisotropic plasmas spontaneously generate small-scale
142: magnetic fields via the Weibel instability (Weibel 1959). Shocks have
143: strong plasma anisotropy in the transition layer separating the
144: upstream and downstream media, as well as in the foreshock where
145: downstream particles escape into the upstream, which provides the free
146: energy for generating magnetic fields with spatial scales on the order
147: of the plasma skin depth, $c/\omega_p$, where $\omega_p$ is the plasma
148: frequency.
149: %These magnetic fields provide the anomalous viscosity, which
150: %mediates a shock, through particles' elastic scattering from the
151: %non-propagating (in the shock frame) magnetic fluctuations (Moiseev
152: %and Sagdeev 1963). Anisotropic flow energy is converted into
153: %isotropic themal energy, as measured in the frame of the downstream
154: %plasma.
155: Medvedev and Loeb (1999) and Gruzinov and Waxman (1999) argued that the relativistic form of this
156: instability (Yoon \& Davidson 1987) could macroscopically magnetize
157: the downsteam plasma where the GRB radiation arises (Larmor radii
158: small compared to flow scale). Through numerical and analytic
159: studies, they and various authors (Silva {\it et al.} 2003, Medvedev
160: {\it et al.} 2005) have argued that the instability forms filaments of
161: electric current and $B$ field. These filaments merge and cause
162: magnetic energy to cascade from the initial microscopic scale $\sim
163: c/\omega_p$ to larger scales. Thus, the filamented plasma becomes
164: magnetized with a $B$ field hypothesized to survive throughout a large
165: fraction of the shocked medium.
166:
167: Such inverse cascades have been observed in 2D and 3D particle-in-cell
168: simulations in the {\it foreshock} region, the part of the shock
169: structure where the upstream and downstream media interpenetrate and
170: the stream filamentation modes grow from noise with most incoming
171: particles still undeflected from the upstream flow.
172: %In the foreshock
173: %region, the currents appear to self-organize into filaments which
174: %merge and generate ever larger scale magnetic structures until the
175: %plasma becomes magnetized.
176: Simulations show that current filaments merge and grow in amplitude
177: until they reach the magnetic trapping limit, where the filament
178: currents and their magnetic fields become comparable to the Alfven
179: limit (Davidson {\it et al.} 1972; Kato 2005; also see Milosavljevic,
180: Nakar, \& Spitkovsky 2006; Milosavljevic \& Nakar 2006a). At that
181: point the particles' orbits on the filament boundaries become chaotic,
182: the filaments disorganize, and scattering from the disorganized
183: magnetic fluctuations halts the streaming of the bulk of the plasma,
184: isotropizing and thermalizing the flow, all within a layer tens to
185: hundreds of skin depths thick (Spitkovsky 2005), in accord with Kato
186: (2005)'s model for shock formation. The magnetic energy is as high as
187: 10-20\% of the bulk plasma flow energy {\it within this scattering
188: layer}, where the density jump between upstream and downstream
189: occurs.
190:
191: %This picture of merging current filaments
192: %has grown into a paradigm for Gamma Ray Burst physics, with regular
193: %invocation of this physical process as the explanation of the magnetic
194: %fields required in synchrotron models (Piran 2005ab, Katz et al 2006).
195:
196: These initially small-scale B-fields must survive for tens of
197: thousands to millions of inverse plasma periods to serve as the source
198: of the magnetization invoked in synchrotron models of GRB emission
199: (Gruzinov \& Waxman 1999; Piran 2005ab; Katz, Keshet, \& Waxman 2007). Long-lived fields are also required in Diffusive Fermi Acceleration (DFA)
200: models used to explain the appearance of the nonthermal particle
201: spectra observed through their synchrotron and inverse Compton
202: emission in GRBs, and in pulsar wind nebulae (PWNs) and other sites of
203: nonthermal photon emission in relativistic flows. However, present
204: simulations (Kazimura {\it et al.} 1998; Silva {\it et al.} 2003; Frederiksen {\it et al.}
205: 2004; Hededal {\it et al.} 2005; Nishikawa {\it et al.} 2003, 2005)
206: have not followed the flow through the shock transition layer into the
207: downstream region because they employ periodic boundary conditions,
208: are too small, or both. The question of the structure and long-term survival of
209: the B-fields has remained open (see for instance, Gruzinov \& Waxman 1999; Gruzinov 2001ab; Medvedev
210: {\it et al.} 2005).
211:
212: In this paper, we show via linear theory that the phase mixing between
213: individual particles and organized macroscopic currents implies
214: rapid decay of the magnetic energy in the downstream medium. We begin
215: by briefly describing the essential features of the plasma produced by
216: the shock via numerical simulations in \S \ref{sec:simulation}. These
217: simulations show the downstream magnetic structures are
218: non-propagating in the frame of the downstream medium, and are intermittent spatially, organized into clumps
219: and filaments of magnetic field with typical diameter $\sim 10-20$ skin depths, immersed into a highly isotropic plasma.
220: %(oriented across the flow direction in the 2D results exhibited here;
221: %3D results, which form large scale loops and arcs, will be discussed
222: %elsewhere)
223: In this downstream isotropic Maxwellian particle distribution, we
224: calculate the damping rates from Vlasov linear response theory,
225: assuming the particles are unmagnetized (Larmor radii $\gg$ magnetic
226: clump diameter), recovering a result due to Mikhailovskii (1979) in \S
227: \ref{sec:theory} and Appendix A. Taking snapshots from the
228: simulations as initial conditions, we calculate the decay of the
229: magnetic field as a function of time and position, and compare these
230: theoretical calculations with simulations. We find that the theory
231: does reasonably well in estimating the decay rate of the total
232: magnetic energy as $t^{-q}$ with $q\sim 1$. However, it overestimates
233: the damping rate of shorter wavelength modes. We speculate in
234: \S\ref{sec:trap} that this discrepancy may result from the magnetic
235: trapping of a large fraction of the particles, which suppresses the
236: disorganizing effect of phase mixing on the currents. We also discuss
237: the effects of an inverse cascade on the persistence of magnetic
238: fields, though we find little evidence for its relevance from
239: numerical simulations. We summarize our results and draw some
240: implications for Fermi acceleration and for the magnetization required
241: for synchrotron emission in models of GRBs and other systems in
242: \S\ref{sec:discussion}.
243:
244:
245: %mediate the
246: %acceleration of a small number of particles to large $\gamma$'s.
247:
248:
249:
250:
251: %These small scale fields have been invoked as In order to do so, these
252:
253:
254:
255:
256:
257: %In
258: %this scenario, these particle radiate through synchrotron emission in the
259: %magnetic fields created at the shock interface.
260:
261: %For this magnetic energy to survive with little or no decline of the
262: %magnetic energy for the scale of the post shock flow that gives rise
263: %to the observed radiation ($10^6 - 10^{10}$ skin depths, Piran 2005)
264: %requires the now disorganized magnetic structures to continue
265: %merging to ever larger length scales. If this is the case, the GRB
266: %magnetization problem would be solved. Instead, simulations to date
267: %have shown that these fields are strongly damped {\it downstream},
268: %although because of residual electromagnetic noise introduced by the
269: %finite (although large) number ($10^{8.5} - 10^{10}$)of macroparticles
270: %employed, the asymptotic state of the magnetic energy remains
271: %inconclusive.
272:
273:
274: \section{Simulation Results}\label{sec:simulation}
275:
276: Spitkovsky (2005) and Spitkovsky and Arons (in prep) describe a series
277: of 2D and 3D simulations of relativistic shock waves in $e^+e^-$
278: plasmas, for varying values of the upstream magnetic field $B_1$,
279: including $B_1 = 0$. These are Particle-in-Cell (PIC) simulations
280: (Birdsall and Langdon 1991), using the code {\it TRISTAN-MP},
281: developed by one of us (AS). It is a heavily modified descendant of
282: the publicly available code {\it TRISTAN} (Buneman 1993). We simulate shocks
283: by injecting cold relativistic plasma particles at one end of a large
284: domain and allowing the particles to reflect off a fixed conducting
285: wall at the other end of the box. In this paper, we present 2D
286: calculations utilizing boxes as large as 50,000 x 2048 cells with up
287: to $1.35\times 10^{10}$ particles which allows us to fully resolve
288: shock formation. Although we track all three velocity and field
289: components, due to the two-dimensional symmetry only particle
290: velocities in the plane of the simulation are non-zero for initially
291: cold flow, and only the out-of-plane component of the magnetic field is
292: excited by in-plane currents. In-plane electrostatic fields are also
293: included.
294:
295: The interaction of the reflected pair plasma with the incoming stream
296: forms a shock, which propagates toward the plasma injection surface.
297: In the simulations, one plasma skin depth spans 10 cells based on the upstream parameters,
298: i.e., $\lambda_{s1} = c/\omega_{p1} = \sqrt{m_\pm c^2 \gamma_1 /4\pi
299: e^2 n_{1}}$, where $n_{1} $ is the upstream total density of
300: electrons and positrons (as measured in the frame of the simulation),
301: $\omega_{p1}$ is the upstream plasma frequency, and $\gamma_1 m_\pm
302: c^2$ is the upstream flow energy/particle. The time step is $\Delta t
303: = 1/20 \omega_{p1}$. In the upstream skin depth units, the largest
304: boxes are then 5000 x 205 $c/\omega_{p1}$. The longest simulation was
305: evolved for $5300 \omega_{p1}^{-1}$. We have checked, by using boxes
306: of increasing transverse dimension, that the periodic boundary
307: conditions used on the transverse coordinate do not affect the scale
308: of the magnetic structures formed. In the results exhibited below, we
309: use $64$ particles per cell ($32$ per species) in order to suppress
310: particle noise, which enables us to follow the dynamics of small
311: amplitude fields. We have performed convergence studies, varying the
312: number $N$ of particles per cell from $4$ to $64$, confirming that the
313: noise level decreases as $1/\sqrt{N}$, while the gross qualities of
314: the shock structure remain the same. Figure \ref{fig:shock} shows the
315: snapshots of density and magnetic energy from a typical 2D simulation.\footnote{Note that we have defined the magnetic energy fraction $\epsilon_B\equiv B^2/4 \pi
316: \gamma_1 n_1 m c^2$ in terms of
317: the {\it upstream} kinetic energy density, instead of the downstream
318: thermal energy density as is a common practice in the GRB community
319: (c.f., Piran 1999).} Coordinates are in units of the upstream skin
320: depth, $c/\omega_{p1}$. In the simulation shown, the upstream flow
321: moves to the left with $\gamma_1 = 15$.
322:
323: \begin{figure*}
324: \plotone{fig1.subsamp.ps}
325: \caption{Snapshot of a region from a large 2D relativistic shock
326: simulation. Incoming $\gamma=15$ flow is moving to the left,
327: while the shock moves to the right at $\approx 0.5c$. The postshock
328: plasma is stationary in the frame of the simulation.
329: a) Density structure in the simulation plane
330: showing the plasma density enhancements in the foreshock region that
331: steadily grow up to the shock transition region, where the density
332: becomes homogeneous. Density is normalized in units of the plasma
333: density far upstream. The density
334: jump, $\langle n_2 \rangle / \langle n_1 \rangle = 3.13$, shown is
335: exactly what is predicted by the hydrodynamic jump conditions for
336: $\gamma_1 = 15$ in a 2D gas. b) Magnetic energy, normalized in terms of upstream energy of the
337: incoming flow: $\epsilon_B=B^2/4 \pi
338: \gamma_1 n_1 m c^2$.
339: The upstream magnetic filaments, which can be visualized as
340: sheets coming out of the page, that are
341: formed by the Weibel instability reach a peak just before the
342: shock. These filaments become clumps of magnetic field, which can be
343: visualized as cross-sections of loops that are transverse to the page in
344: the downstream region, where they slowly decay away. Note that
345: the B-field, which is organized into upstream filaments
346: or downstream clumps, always points in or out of the page.
347: A power law
348: scaling, $\epsilon_B^{1/4}$, was applied to stretch the color table
349: to show weak field regions; this is reflected in the colorbar. c)
350: Plasma density averaged in the transverse direction as a function of
351: the distance along the flow. d) Magnetic energy density averaged in
352: the transverse direction, as a function of distance along the flow.}
353: \label{fig:shock}
354: \end{figure*}
355:
356: Our simulations are large enough to permit the complete development of
357: the shock and show the main features of contemporary collisionless
358: shock simulations even in reduced dimensionality. For instance, we
359: see the factor of $\approx 3.13$ increase in density between the upstream and
360: the downstream (Fig. \ref{fig:shock}c), which is the expected compression
361: factor for this 2D plasma when the plasma properties are measured in
362: the rest frame of the downstream plasma (Gallant {\it et al.} 1992;
363: Spitkovsky and Arons, in prep). Current filaments (geometrically,
364: these are out-of-plane sheets) show up as the enhancement in plasma
365: density and magnetic energy density in the foreshock (Fig.
366: \ref{fig:shock}ab). The scale of the filaments grows towards the shock
367: through merging. The shock is located where the density filaments
368: completely merge and are replaced by a quasi-homogeneous medium. The
369: subject of the saturation of the Weibel instability will be explored
370: in greater detail by Spitkovsky and Arons (in prep).
371:
372: %occurs when
373: %the
374: %currents reach a critical strength set by the Alfven current limit (Kato 2005;
375: %Davidson {\it et al} 1972),
376:
377: %fields
378:
379:
380: %The free particle streaming in the filaments ends when the transverse
381: %magnetic fields in the filaments grow large enough to deflect
382: %particles by 90 degrees, a condition known as magnetic trapping, which
383: %has been long identified as the saturation mechanism for the
384: %non-relativistic Weibel instability (Davidson {\it et al} 1972).
385: %Magnetic trapping is equivalent to the Alfven critical current
386: %condition for saturation when the densities of the two
387: %counter-streaming beams of particles are the same.
388:
389: At the shock transition layer, the filaments disorganize and become
390: clumps of magnetic energy (in 2D the only appreciable magnetic
391: component is the out of plane $B_z$). Note that these magnetic clumps
392: lose intensity the further downstream they are from the shock (Figure
393: \ref{fig:shock}b). The magnetic energy peaks before the density
394: completes its rise (as we see in comparing c and d in Fig.
395: \ref{fig:shock}), i.e., the instability saturates at the Alfven
396: critical current before the shock fully develops. We also note that
397: the particle distribution function changes from an anisotropic
398: free-streaming population to an isotropic (in the downstream rest
399: frame) thermal population. In the downstream medium, we find that the
400: difference between the perpendicular and parallel momentum is
401: extremely small, $<1\%$.
402:
403: As Figures \ref{fig:shock} b and d suggest, even though locally in the
404: shock magnetic energy can reach close to equipartition ($epsilon_B \approx 15\%$ when
405: averaged over the transverse dimension), the magnetic fields decay in
406: the downstream region of the shock. To study this in further detail,
407: we present several cross sections of the plasma at times separated by
408: 450 $\omega_{p1}^{-1}$ in Figure \ref{fig:shock later}. Since our
409: simulation is performed in the downstream frame (frame of
410: the reflecting wall), we see the shock propagate through the box at
411: $\approx 0.5 c$ (value appropriate for 2D relativistic gas).
412:
413: Downstream from the shock, the magnetic clumps weaken as a function of time
414: and appear to be non-propagating. In Figure \ref{fig:shock later}, we see
415: this in panels a-d and in the black through blue curves in panel e,
416: which show magnetic energy averaged over the transverse dimension of
417: the simulation at times from panels a-d. Although the separation
418: between prominent clumps that have larger field strength seems to
419: increase with time, this is due to the faster disappearance of
420: small-scale clumps, presumably due to decay, while the location or
421: size of strong clumps does not significantly evolve. There is little
422: evidence for clump merging far from the shock. In Figure \ref{fig:shock later}
423: we also plot a vertical line showing the location where we will
424: decompose the magnetic fields into Fourier modes later in \S3, in
425: order to study the comparison of the theory of B-field decay with the
426: numerical experiments.
427:
428: \begin{figure*}
429: \plotone{fig2.subsamp.ps}
430: \caption{ Cross-sectional views of the magnetic energy density for the
431: same region as Figure \ref{fig:shock}, but at different times
432: separated by 450 $\omega_p^{-1}$ (a-d). The time of Figure
433: \ref{fig:shock} corresponds to panel b) in this plot. In panel e) we
434: plot the B-field energy, averaged over y-axis as a function of
435: distance along the flow direction with one curve for each of the top panels: (a)
436: black, (b) red, (c) green, (d) blue. We also plot a vertical line
437: at $x = 840 c/\omega_p$ to define the region over which we perform a
438: modal analysis in \S\ref{sec:theory} and Figure
439: \ref{fig:fftsingle}.}
440: \label{fig:shock later}
441: \end{figure*}
442:
443: %There is little evidence
444: %of merging of the downstream magnetic structures, which now take the
445: %form of quasi-cylindrical $\theta$ pinches, with B organized in
446: %filaments directed out of the plane of the figure.
447:
448: Finally, we mention that 3D shock simulations also show similar
449: behavior (including the lack of substantial downstream merging of magnetic
450: structures) for the shorter times that can be followed with present 3D
451: simulations. Topologically, the magnetic clumps in 2D become large
452: looping structures in 3D. If our 2D plane is thought of as a slice
453: through a 3D simulation, the 3D loops connect field emerging
454: from one clump and returning to another. The orientation of 3D loops
455: would be mostly perpendicular to the original direction of the flow.
456: The particle distribution function is also an isotropic, relativistic
457: Maxwellian in the downstream region.
458:
459: %3D simulations show similar behavior (including lack of substantial
460: %downstream merging of magnetic structures), for the rather shorter
461: %times that can be followed. The filaments with straight magnetic
462: %fields form large looping structures with radii of curvature large
463: %compared to filament radius, as is illustrated in figure
464: %\ref{fig:threeDB}.
465: %
466: %\begin{figure}[H]
467: %\plotone{bfield.png}
468: %\caption{Visualization of the magnetic energy density in a 3D simulation, showing the upstream filaments along the flow to the right and the downstream loops, decaying
469: %with distance from the shock front. }
470: %\label{fig:threeDB}
471: %\end{figure}
472:
473:
474: \section{Downstream Evolution of Magnetic Turbulence}\label{sec:theory}
475:
476: The simulations summarized in \S\ref{sec:simulation} suggest that
477: magnetic turbulence decays in the isotropic post-shock plasma. We now
478: attempt to understand this theoretically with the goal of finding a model that
479: allows extrapolation of the magnetic evolution beyond the length of time that can be
480: studied in direct shock simulations. The simulations show that the downstream
481: plasma is isotropic and the downstream particle distribution
482: function is well described by a relativistic Maxwellian. For the
483: purposes of this section, we {\it assume} the downstream field
484: amplitudes are so small that magnetic trapping is unimportant for
485: almost all particles; therefore, their orbits are almost straight
486: lines. We will revisit this assumption in \S\ref{sec:trap}.
487:
488: We begin by deriving the linear plasma response for electromagnetic
489: fluctuations in an isotropic relativistic plasma with small field
490: fluctuations (Mikhailovskii 1979). The Vlasov equation for each species is
491: \begin{equation}\label{eq:linearize vlasov}
492: \frac {\partial f_s}{\partial t} + {\boldsymbol v} \cdot {\boldsymbol \nabla} f_s +
493: \frac {q_s}{m_e} \left({\boldsymbol E} + \frac {{\boldsymbol v}} {c} \times {\boldsymbol B} \right)
494: \cdot{\boldsymbol \nabla}_p f_s = 0.
495: \end{equation}
496: Here, $s$ is the species label, $+$ for positrons and $-$ for electrons,
497: with $q_{\pm} = \pm e$. We linearize this equation with $f_s
498: \rightarrow f_{0s} + \delta f_s$, ${\boldsymbol B} \rightarrow \delta
499: {\boldsymbol B}$ and ${\boldsymbol E} \rightarrow \delta {\boldsymbol
500: E}$, with the initial conditions downstream of the shock that tell
501: us that $\delta E$ and $\delta B$ are small in the sense that $(\delta
502: E^2 + \delta B^2)/8\pi nT \ll 1$. Then $\delta f_s /f_{0s}$ is also
503: small. The linearized Vlasov equation is
504: \begin{equation}\label{eq:vlasov}
505: \frac {\partial\delta f_s}{\partial t} + {\boldsymbol v} \cdot {\boldsymbol \nabla}\delta f_s
506: + \frac {q_s}{m_e} \left(\delta {\boldsymbol E} + \frac {{\boldsymbol v}} {c} \times \delta {\boldsymbol B} \right)\cdot{\boldsymbol \nabla}_p f_{0s} = 0.
507: \end{equation}
508: The plasma couples to the field through the Maxwell equations
509: \begin{eqnarray}\label{eq:maxwell}
510: {\boldsymbol \nabla} \times \delta {\boldsymbol E} &=& -\frac 1 c \frac {\partial \delta {\boldsymbol B}} {\partial t},\\
511: {\boldsymbol \nabla} \times \delta {\boldsymbol B} &=& \frac {4\pi} {c} \delta {\boldsymbol j} + \frac{1}{ c} \frac {\partial \delta {\boldsymbol E}} {\partial t},\label{eq:maxwell2}
512: \end{eqnarray}
513: where $\delta {\boldsymbol j} = \sum_s q_s\int {\boldsymbol v} \delta f_s d^3 p$ is the current density.
514:
515: We orient coordinates so that $\delta {\boldsymbol E}$ is along $x$, the wave vector lies
516: along $y$, and $\delta {\boldsymbol B}$ is along $z$. Assuming a wave solution with
517: amplitudes $\propto \exp\left(-i\omega t + i k y\right)$, equation
518: (\ref{eq:vlasov}) becomes
519: \begin{equation}
520: i \left(kv_y - \omega\right)\delta f_s + \frac {q_s}{m_e} \left(\delta E +
521: \frac {v_y} c \delta B\right)\frac{\partial f_{0s}}{\partial p_x} -
522: \frac {q_s}{m_e} \frac {v_x} {c} \delta B \frac {\partial f_{0s}}{\partial p_y} = 0.
523: \end{equation}
524: Using equation (\ref{eq:maxwell}) to eliminate $\delta E$ yields
525: \begin{equation}\label{eq:deltaf}
526: \delta f_s = i \frac {q_s \delta B}{m_ekc} \left(\frac {\partial f_{0s}} {\partial p_x}
527: + \frac {kv_x}{\omega - kv_y} \frac {\partial f_{0s}}{\partial p_y} \right).
528: \end{equation}
529: Using (\ref{eq:maxwell2}) yields
530: \begin{equation}\label{eq:deltaB}
531: i\left(\omega^2 - k^2 c^2\right)\delta B = -4\pi kc \delta j.
532: \end{equation}
533: Using (\ref{eq:deltaf}) in the definition of $\delta j$ yields
534: \begin{equation}
535: \delta j = -i \chi\omega \delta E
536: \label{eq:j2E}
537: \end{equation}
538: where $\chi$ is the plasma susceptibility,
539: \begin{equation}
540: 4 \pi \chi = \sum_s \frac {\omega_{p,{\rm NR},s}^2}{\omega^2}
541: \int v_x \left(\frac {\partial} {\partial p_x} + \frac {kv_x}{\omega - kv_y}
542: \frac {\partial}{\partial p_y}\right)\frac {f_{0s}}{n_s} d^3p.
543: \label{eq:susceptibility}
544: \end{equation}
545: %and the
546: %dispersion relation
547: %\begin{equation}\label{eq:dispersion}
548: % \frac {k^2 c^2}{\omega^2} - 1 - \frac {\omega_p^2}{4\pi\omega^2}
549: % \int v_y \left(\frac {d} {dp_y} + \frac {kv_y}{\omega - kv_x}
550: % \frac {d}{dp_x}\right)F_0 d^3p = 0,
551: %\end{equation}
552: Here $n_s$ is the number density of the electrons ($s = -$) or
553: positrons ($s=+$) in the downstream plasma, $\omega_{p,{\rm NR},s} =
554: \sqrt{4\pi q_s^2 n_s/m_e}$ is the non-relativistic plasma frequency.
555: We will assume by charge neutrality that in the downstream plasma
556: $f_{0+} = f_{0-}$ and $n_+ = n_-$ so that the sum over equal mass
557: species is trivial.
558:
559: The dispersion relation for normal modes in the plasma plus electromagnetic
560: field is
561: \begin{equation}\label{eq:general_dispersion}
562: \frac{k^2c^2}{\omega^2} - (1 + 4 \pi \chi) = 0.
563: \end{equation}
564: However, we emphasize that the linear relation between the currents and
565: the electric field (eq.[\ref{eq:j2E}]) through the susceptibility (\ref{eq:susceptibility}) applies to all
566: (small amplitude) fluctuations, not only to normal modes.
567:
568:
569: %Since $F_0$ is isotropic, we perform the angular integration to find
570: %\begin{equation}\label{eq:worked out dispersion}
571: % \frac {k^2 c^2}{\omega^2} - 1 - \frac {\omega_p^2}{2\omega^2} \int v
572: % \left[\left(\frac {\omega} {kv}\right)^2 - \frac{\omega}{kv}\left(1
573: % - \left(\frac {\omega} {kv}\right)^2\right)\left\{\frac 1 4
574: % \log\left(\frac {1 + \omega/kv}{1 - \omega/kv}\right)^2 - i\frac {\pi}
575: % 2 \theta\left(\left(\frac {\omega}{kv}\right)^2 - 1\right)\right\}\right]
576: % \frac {dF_0}{dp} p^2dp = 0
577: %\end{equation}
578:
579: We evaluate equation (\ref{eq:susceptibility}) for distribution
580: functions that are isotropic in two and three dimensions in Appendix
581: A. For a two-dimensional isotropic distribution function, we write
582: $f_{0s}({\boldsymbol p}) = f(p_{2d})g(p_z)$, where $p_{2d} =
583: \sqrt{p_x^2 + p_y^2}$ and perform the integral over $p_z$ such that
584: $\int g(p_z) dp_z = 1$. For a three-dimensional isotropic
585: distribution function, we write $f_{0s}({\boldsymbol p}) =
586: f_{0s}(p_{3d})$, where $p_{3d} = \sqrt{p_x^2 + p_y^2 + p_z^2}$. In the
587: three-dimensional case, we confirm Mikhailovskii's (1979) result for a
588: relativistic isotropic plasma. As we show explicitly in equation
589: (\ref{eq:worked out suscept}), subluminal waves ($\omega_r < kc$) are
590: damped, where $\omega_r = \Re(\omega)$. The basic physics is that of
591: Landau damping (Stix 1992) or more precisely, phase mixing.
592: %The standard picture of Landau damping is that there is net power
593: %transfer between a Fourier mode and a particle when the Doppler
594: %shifted frequency vanishes, i.e., $\omega_r - {\boldsymbol k} \cdot
595: %{\boldsymbol v} = \omega_r - k v \cos \phi = 0$.
596: For a subluminous wave in an isotropic relativistic plasma with
597: $\omega_r /k < c$, there always exists some angle $\phi$ between the
598: wave vector $k$ and a particle's momentum where a particle moving with
599: speed $c\beta$ is in phase with the field component, i.e.,
600: $\beta\cos\phi = \omega_r/kc$.
601:
602: When the phase velocity of the field fluctuations is very small
603: compared to the thermal speeds of the particles (which includes the
604: zero wave velocity case, $\Re (\omega) = 0$), Landau damping takes on
605: the character of simple phase-mixing. The currents, which support the
606: field fluctuations, are composed of particles. These current carrying
607: particles have random motions which carry them out of the magnetic
608: structure, which the currents initially support. As a result, these
609: currents and their magnetic field fluctuations are disorganized (or
610: damped) on the transit times of the particles across the field
611: fluctations. It should be emphasized that the formal theory includes
612: both this phase mixing limit of the damping process and the limit in
613: which the phase 4-speeds $\beta_p /\sqrt{1-\beta_p^2}, \; (\beta_p
614: \equiv \Re \omega /kc),$ of the fields' Fourier components are high
615: compared to the particles' random 4-speeds $\beta /\sqrt{1 -\beta^2}$.
616: The applicable limit depends on whether the wave phase 4-velocity is
617: large or small compared to the mean 4-velocity of the particles in the
618: distribution function. Hammett, Dorland \& Perkins (1992) present a
619: simple picture of this process in the case of non-relativistic
620: electrostatic waves. Arons, Norman \& Max (1977) show how phase
621: mixing works for electromagnetic waves in a relativistic Maxwellian
622: plasma. In any case, the end result is that there is a net power flow
623: from fields to particles, {\it i.e.} the fields decay, which we will
624: demonstrate explicitly below.
625: % Hence as long as population of particle are monotonically decreasing with larger momentum ($dF/dp < 0$) then wave energy is lost. An inverted population ($dF/dp > 0$) may add energy to the wave, but such a study is beyond the scope of this analysis.
626:
627: We now study the action of such a plasma on an arbitary field of
628: initial B-field pertubations as generated by the Weibel
629: mediated shock. We evaluate equation (\ref{eq:susceptibility})
630: numerically for two and three dimensions setting $\omega_r = 0$,
631: because of the non-propagating nature of the magnetic clumps, which we
632: infer from visual inspection of the simulations. In the limit $k \ll
633: \omega_p/c$, we find:
634: \begin{equation}\label{eq:chi_limit}
635: 4\pi\chi \approx \left\{\begin{array}{ll}
636: i \frac {\omega_{p}^2}{|k|c\omega} & \textrm{2D} \\
637: i \frac {\pi} 4 \frac {\omega_{p}^2}{|k|c\omega} & \textrm{3D}
638: \end{array}\right. .
639: \end{equation}
640: Note that the 2D and 3D results only vary by a numerical factor.
641: Hence, long wavelength modes have the same qualitative behavior in two
642: and three dimensions.
643: %As we show in Appendix A, this simple form recovers the
644: %exact numerical solution for $\omega_r = 0$. Hence, we will use
645: %equation (\ref{eq:chi_limit}) in the rest of the paper.
646:
647: The plasma susceptibility, which we calculate in Appendix A (see also
648: eq.[\ref{eq:chi_limit}]), can be utilized to calculate the evolution,
649: i.e., the damping or growth, of an initial field of fluctuations. Appendix B
650: contains this calculation in more detail; the result is
651: \begin{equation}\label{eq:damping}
652: \frac {d|\delta B_k|^2}{dt} = -2\gamma_k |\delta B_k|^2,
653: \end{equation}
654: where $\gamma_k = \left(kc\right)^2\omega^{-1} \Im\left(4\pi\chi\right)^{-1}
655: $ (see also eq.[\ref{eq:app_general gamma}]). The
656: asymptotic forms of $\chi$ from equation (\ref{eq:chi_limit}) for 3D
657: and 2D gives (see also eq.[\ref{eq:app_gamma}])
658: \begin{equation}\label{eq:gamma}
659: \gamma_{k} = \left\{\begin{array}{ll} \frac {|kc|^3} {\omega_{p}^2} & \textrm{2D}
660: \\ \frac {4} {\pi} \frac{|kc|^3} {\omega_{p}^2}& \textrm{3D}
661: \end{array}\right. .
662: \end{equation}
663:
664:
665: %Linear theory predicts that short wavelengths are
666: %strongly damped, but long wavelengths persist for longer.
667: We now compare the expectations from linear response theory to the
668: numerical simulations. We begin by taking the Fourier transform of
669: $\delta B$ from our 2D numerical simulations from a downstream region
670: where the shock is fully developed, {\it i.e.}, behind the shock
671: front at $ x = x_0$,
672: \begin{equation}\label{eq:fourier transform}
673: \delta B(x_0,y) = \frac 1 {\sqrt{2\pi}} \int dk_y \delta B_{k_y}(x_0)
674: \exp\left(ik_y y\right)
675: \end{equation}
676: We evolve $\delta B_{k_y}$ in accordance to equation
677: (\ref{eq:damping}) using the asymptotic forms in equation
678: (\ref{eq:gamma}) and compare our analytically evolved spectra with the
679: numerical simulation at later times. In Figure \ref{fig:fftsingle},
680: we take the initial data from a region at $x_0=840 c/\omega_p$, which
681: we marked with a line in Figure \ref{fig:shock later}. To accumulate
682: sufficient statistics, we average the fields over $14\, c/\omega_p$ in
683: the flow direction. We evolve these spectra for 450 (red), 900
684: (green), and 1350 $\omega_p^{-1}$ (blue) using equation
685: (\ref{eq:gamma}) and compare this to snapshots taken from our
686: numerical simulations at these times. Theory and simulation agree at
687: very low wavenumber ($k_yc/\omega_p \lesssim 0.2$). However, theory
688: overpredicts the cutoff in power at larger k. The discrepancy
689: suggests that linear theory is insufficient to describe the nature of
690: downstream magnetic turbulence and that additional physics is needed
691: (see \S\ref{sec:trap}).
692:
693: \begin{figure}
694: \plotone{f3.eps}
695: \caption{ Spectral evolution of magnetic field from the slice at $840 c/\omega_p$ from Fig. \ref{fig:shock later}. Initial field spectrum (black solid line) is plotted after $450\,\omega_p^{-1}$ (red),
696: $900\,\omega_p^{-1}$ (green), and $1350\,\omega_p^{-1}$ (blue) based on simulation data. Dashed curves represent analytic evolution of the initial field and overpredict decay of short-wavelength structures.
697: %Comparison of analytic evolution (dashed line) with the 2D
698: %simulation (solid lines) at $450\,\omega_p^{-1}$ (red),
699: %$900\,\omega_p^{-1}$ (green), and $1350\,\omega_p^{-1}$ (blue) of an
700: %initial field taken from a region behind the shock. The spectral decomposition of the initial
701: %field (black solid line) is also shown.
702: \label{fig:fftsingle} }
703: \end{figure}
704:
705: The lack of power in short wavelength modes suggests that the total
706: B-field is determined by long wavelength modes. We now use equation
707: (\ref{eq:damping}) to find a simple decay law for the total B-field:
708: \begin{equation}
709: \label{eq:fourier evolution}
710: \delta B(x_0,y,t) = \frac 1 {\sqrt{2\pi}} \int dk_y \delta B_{k_y}(x_0)
711: \exp\left(ik_y y - \gamma_k t\right),
712: \end{equation}
713: where $\delta B_{k_y}(x)$ is defined in equation (\ref{eq:fourier
714: transform}). Inserting equation (\ref{eq:gamma}) into (\ref{eq:fourier evolution}) and
715: integrating, with $\delta B_{k_y} =a k_y^p$, where $p$ is the
716: long wavelength spectral index and $a$ normalizes the amplitude, we find
717: \begin{eqnarray}
718: \delta B(x_0,y,t) &=& \frac 1 {\sqrt{2\pi}} a \int_{k_0}^\infty dk_y k_y^p
719: \exp\left(-\alpha \omega_{p}t \left(\frac {k_yc}
720: {\omega_{p}}\right)^3\right) \\
721: &\approx& \frac{1}{3 \sqrt{2\pi}} a \left( \frac{\omega_{p}}{c} \right)^{p+1}
722: \Gamma\left(\frac {p+1}{3} \right)
723: \left(\frac{1}{ \alpha \omega_{p} t } \right)^{(p+1)/3}
724: \label{eq:decay-power}
725: \end{eqnarray}
726: where we have taken $y=0$ without loss of generality, $\alpha = 4/\pi$
727: in 3D and $\alpha = 1$ (2D), $k_0 $ is small such that $\omega_p t
728: (k_0 c/\omega_p)^3 \approx 0$, and $\Gamma$ is the gamma function.
729: Note that $\omega_p$ is for the downstream plasma frequency. However, this
730: detail is irrelevant in terms of determining the power law as a
731: function of $t$.
732: % - in the simulations, where time is
733: %measured in units of $\omega_{p1}^{-1}$ and length in units of
734: %$c/\omega_{p1}$, the transverse size is $200 c/\omega_{p1}$ in the 2D
735: %models.
736: Setting $k_0$ to be small is a safe approximation in the simulations,
737: and of course is excellent in much larger astrophysical systems.
738: Thus if the initial spatial spectrum is a power law in wave number
739: $|\delta B_k|^2 \propto k^{2p}$, then the theory predicts $\delta B^2
740: \propto t^{-2(p+1)/3}$.
741:
742: %Magnetic field energy appears initially to follow this trend in our
743: %numerical simulations as shown in .
744: We study the region immediately following the peak of magnetic energy
745: in the shock front (where it reaches $\delta B^2 = B_{\rm max}^2$) at
746: $x_{\rm peak}$ and plot its value as a function of position in the
747: postshock region. For a shock moving at constant velocity, we have
748: $x_{\rm peak} - x \propto t$. Hence $\delta B^2 \propto (x_{\rm peak}
749: - x)^{-2(p+1)/3}$. Our numerical simulations are extremely suggestive
750: that the magnetic energy density follows the $p = 0, \; \epsilon_B =
751: \delta B^2/8\pi \propto t^{-2/3}$ decay expected for an initially flat
752: magnetic spectrum at early times, then steepens to a $t^{-1}$ decay at
753: later times as shown in Figure \ref{fig:eb}, until the noise finally
754: swamps the signal.\footnote{Gruzinov (2001a) performed shock
755: simulations (initiated by a collision between two $e^\pm$ plasmas),
756: which show decay of the magnetic energy averaged over the dimension
757: across the flow similar to the early phases of the decay we report
758: here.}
759: This $t^{-1}$ decay rate would imply an initial spatial
760: index of $p=1/2$ at low wavenumber, though our analysis of additional
761: simulations with a large transverse spatial scale suggest $p=0$. The
762: difference in the index of the decay law expected from the theory and measured from
763: the simulations may be due to magnetic trapping (see \S\ref{sec:trap}).
764: Gruzinov (2001b) offered an alternative
765: explanation of $t^{-1}$ decay of the magnetic energy density observed
766: in 2D simulations of Weibel instability in counterpropagating pair
767: plasmas. In that simulation the beams were moving perpendicular to
768: the simulation plane\footnote{Being orthogonal to the direction of
769: motion these simulations did not form a shock and retained some
770: counterstreaming at late times. In contrast, our simulations lose
771: the counterstreaming once the shock forms.}. In order to explain
772: the field decay, Gruzinov (2001b) had to assume that the magnetic
773: structures increase their size at the Alfven velocity. In our shock
774: simulations, the downstream magnetic structures do not significantly
775: expand (\S2), hence the similarity of decay law between the two
776: simulations is likely a coincidence. The picture of filament expansion
777: and merging of Gruzinov (2001b) is likely valid very close to the
778: shock, but is not observed in the longer evolution of the downstream
779: plasma.
780: %However, his argument assumed that the expanding
781: %magnetic structures, which he observed in his colliding beam
782: %simulations,\footnote{Gruzinov (2001b) simulates the evolution of
783: % magnetic field during Weibel instability driven by collision of two
784: % infinite plasma streams. The plane of the 2D simulation is
785: % perpendicular to the direction of motion of the streams. As such,
786: % this simulation does not form a shock. In contrast, our simulation
787: % is in the plane along the motion of the streams, and allows density
788: % compression. } increase their size at the Alfven velocity. As we
789: %have pointed out in \S2, these magnetic structures neither propragate
790: %nor significantly expand. Hence, Gruzinov's reasoning is not
791: %applicable to our results.
792: %We
793: %discuss Gruzinov's (2001) alternative explanation of $t^{-1} $ decay
794: %of the magnetic energy density, based on conservation of the vector
795: %potential (which is a much more severe requirement than conservation
796: %of canonical momentum), in Appendix \ref{sec:canonical momentum}.
797:
798: \begin{figure}
799: \plotone{f4.ps}
800: \caption{Magnetic energy density (in units of upstream kinetic energy) as a function of position downstream of the shock. A broken power law proportional $(x-x_{\rm
801: peak})^{-2/3}$ fits well at early times, but a $(x-x_{\rm
802: peak})^{-1}$ power law fits better at later times. }
803: \label{fig:eb}
804: \end{figure}
805:
806:
807:
808: %However, the 2D spectra are rather noisy, since they are 1D spectra over a single
809: %transverse dimension. Spectra from 3D simulations are rather better behaved,
810: %since they are taken over a 2D y-z plane, as shown in Figure \ref{fig:fftpeak}
811: %These spectra reveal an interesting departure of the theory from the experiments,
812: %as shown in Figure \ref{fig:fftpeak}
813: %\begin{figure}[H]
814: % \epsscale{1.0} \plotone{manyBkwithfitbw.png}
815: % \caption{Decay of the magnetic fluctuation power spectra from a 3D simulation
816: % run with $512 \times 512 \times 6000$ cells, and 4 particles per cell initially.
817: % The four curves show the spectra with time increasing from top to bottom. Each curve
818: % is separated by $\Delta t = 182/\omega_{p1} $. The fitted straight lines are
819: % $| \delta B_k|^2 \propto \exp[- g \omega_{p1}t (kc/\omega_{p1}) ]$ - the decay
820: % is better fit with a damping rate linear rather than cubic in $k$. In addition, these
821: % 3D spectra show substantial magnetic merging at early times, shown by the move of
822: % the magnetic energy peak to smaller wave number. This merging dies away as the
823: % decaying magnetic clumps become more spatially isolated. }
824: %\label{fig:fftpeak}
825: %\end{figure}
826: \begin{figure*}
827: \plotone{fig5.subsamp.ps}
828: \caption{Segments of particle orbits from the high resolution 2D simulation overplotted on a snapshot of magnetic energy downstream of the shock. Typical sizes of magnetic clumps are
829: 10-20 $c/\omega_p$. The shock is near the right boundary of this slice, moving to the right. The large angle deflections in particle orbits suggest the Larmor
830: radii of many particles in this slice %$\sim 300 c/\omega_{p1}$
831: are of the same order as the sizes of the magnetic clumps.
832: \label{fig:trapped}}
833: \end{figure*}
834:
835: \section{Magnetic Trapping}\label{sec:trap}
836:
837: Our simulations and theory suggest power law $t^{-q}$ with
838: $q \sim 1$ temporal decay of the total magnetic energy density in
839: rough agreement with each other. However, linear theory and numerical
840: simulations disagree on the decay rate for short wavelengths, which
841: hinders a determination of the ultimate fate of the fields on times
842: longer than several thousand plasma periods. This discrepancy may
843: arise from the nonlinear effects of magnetic trapping. Particles in
844: the magnetic fields do not follow straight line trajectories that are
845: weakly perturbed, but are partially trapped and strongly deflected.
846: We illustrate this point from following test particles' orbits in
847: numerical simulations in Figure \ref{fig:trapped}. The test particles
848: suffer large deflections from straight-line orbits as they encounter
849: magnetic clumps. Therefore, the Larmor radii of many of the test
850: particles are of the same order of the sizes of these clumps or
851: smaller. Indeed closer inspection of some of the test particle orbits
852: that are embedded inside the clumps suggests that they are completely
853: trapped.
854:
855:
856:
857:
858: %To hammer this point further, we plot the distribution of Larmor radii
859: %in the whole downstream region in Figure \ref{fig:larmor}. Since the
860: %rough size scale of these magnetic clumps are 50-100 cells, a
861: %substantial number of particles are trapped in these clumps.
862: %\begin{figure}
863: %\plotone{LarmorRadiiDistrLinear.png}
864: %\caption{Distribution of downstream particles' formal Larmor radii
865: % $m_\pm c^2 \gamma/eB $. Since the scale of the magnetic structure is
866: % typically 50-100 cells, a substantial fraction of particles are likely trapped.
867: % \label{fig:larmor}}
868: %\end{figure}
869:
870: These strong departures from weakly perturbed particle dynamics may be
871: the cause of the decreased damping at large wavenumber found in the
872: simulations, compared to the predictions of unmagnetized plasma theory.
873: Magnetic trapping may also modify the decay of magnetic fields at
874: small wavenumber, leading perhaps to a different decay law instead of
875: expected $t^{2/3}$ decay law that we find from our simple linear theory
876: (eq.[\ref{eq:decay-power}] with $p=0$).
877: The temporary (and permanent, for some particles)
878: binding of particles to the spatially intermittent magnetic fields
879: reduces the effect of rapid phase mixing central to the unmagnetized
880: plasma damping theory. Magnetic trapping already plays a critical
881: role in the saturation of the initial Weibel instability at the shock
882: transition region (Kato 2005; Davidson {\it et al.} 1972). Our
883: analysis suggests that trapping also plays an important role
884: downstream even though the {\it average} magnetic amplitudes are
885: greatly reduced from the shock transition region - the essential point
886: is that within the isolated filaments, the magnetic pressure is not
887: small. The problem of the damping of isolated magnetic structures in
888: a plasma with partial magnetization illustrated in Figure
889: \ref{fig:trapped} will be the subject of a separate investigation.
890:
891:
892: \section{Discussion}\label{sec:discussion}
893:
894: We have studied the downstream evolution of magnetic turbulence in the
895: context of a collisionless $e^+e^-$ shock both analytically and
896: numerically. Our large scale 2D simulations show the formation of
897: filaments in the foreshock region which merge and grow
898: until they reach the shock transition region. Past the shock
899: transition region, these filaments break up into magnetic clumps in a
900: quasi-homogenous medium where the background particle distribution
901: function is an isotropic Maxwellian. In such a background, we showed
902: that magnetic energy will decay like $t^{-q}$ with $q\sim
903: 1$ due to untrapped particle phase mixing, which is broadly consistent
904: with our numerical simulations. In detail, the theoretical decay
905: rates depend more strongly on wavelength than is seen in the numerical
906: experiments. We suggested that magnetic trapping may play an
907: important role in resolving this discrepancy. Trapping can lead to
908: MHD-like behavior and possible long time persistence of some of the
909: magnetic energy in spatially intermittent, more strongly magnetized
910: subregions. This effect is a subject of further study.
911:
912: % In addition, a
913: %sufficiently fast inverse cascade may reduce magnetic decay. However,
914: %our simulations show no evidence for such a mechanism in the
915: %downstream medium.
916:
917: Magnetic field decay may be alleviated via an inverse cascade from
918: small scales to large scales. An inverse cascade via current filament
919: merging operates strongly in the {\it foreshock} region (Silva {\it et
920: al.} 2003; Spitkovsky 2005). However, it is unclear if this
921: picture can be applied to the downstream region. The magnetic
922: filaments so prevalent in the upstream region are gone and are
923: replaced by isolated magnetic clumps (loops), a result common to both 2D and
924: 3D simulations. Filament merging does occur, but is confined to the
925: foreshock, where the filaments exist. We see little evidence of
926: merging magnetic clumps. This is reinforced by Katz {\it et al.} (2007), who
927: suggest via self-similar arguments that the inverse cascade does not
928: operate downstream of the shock. Hence, the magnetic energy should damp
929: away in the downstream region.
930:
931: We have not studied in detail how these magnetic clumps are confined
932: in the downstream plasma. However, Fujita et al. (2006) have found
933: these same magnetic clumps in simulation of the non-relativistic
934: Weibel instability that is appropriate for galaxy clusters. In this
935: case, they argue that pressure of the external medium confines these
936: magnetic clumps. The strong perturbation from straight line motion,
937: which we studied in \S\ref{sec:trap} for test particles, suggest that
938: the magnetic clumps in our simulation are also sensitive to momentum
939: transfer between particles and itself. That is, the magnetic clumps
940: in our simulations are also confined by external pressure.
941:
942: Whatever the final fate of the spatially intermittent fields, our
943: study of the nature of $e^+e^-$ shocks suggests that the belief in the
944: persistence of Weibel generated magnetic fields with strengths
945: comparable to those appearing within the shock transition is overly
946: optimistic. Magnetic field energies tend to decline rapidly after
947: about a few hundred plasma skin depths. Only the very long term
948: evolution is still open to some question, i.e., does $\langle
949: \epsilon_B \rangle$ settle at some value smaller than $10^{-3}$, or
950: does the spatially intermittent magnetic field decline to zero?
951:
952: The rapid decay suggested by our analysis puts severe constraints on
953: the synchrotron emission mechanisms for GRBs. The width of the
954: emitting region is $\sim 10^9 c/\omega_p$ (Piran 2005b), which is much
955: larger than the region over which we expect magnetic fields to
956: persist. However, decaying magnetic fields may not be inconsistent
957: with GRB observations. Pe'er \& Zhang (2006) suggest that the prompt
958: emission from the internal shock may be more consistent with a
959: decaying magnetic field component rather than a persistent field
960: component. A field that persists over a scale of $10^4-10^5$ plasma
961: skin depths fits the spectra better at low-energies than a
962: field that persists over the entire thickness of the shell ($10^9$
963: plasma skin depths). Small magnetized regions may also be important
964: in the context of the afterglow (Rossi \& Rees 2003).
965:
966: In young pulsar wind nebulae, post
967: shock magnetic fields averaged over the whole latitudinal extent of
968: the observed emission tori have energy densities of a few
969: percent of the post shock plasma energy density. It is possible that
970: weaker magnetic fields exist near the midplane of the equatorial flow, a
971: region of particular interest to the conversion of flow energy into
972: the observed nonthermally emitting spectra of $e^\pm$. If so, the
973: Weibel mediated shock dynamics studied here may be of relevance to
974: these systems' behavior.
975:
976: It is also possible that weak systematic upstream magnetic fields are
977: of essential importance, and that shocks in completely unmagnetized
978: plasmas are an oversimplification. Suprathermal particle generated at
979: the relativistic shock front may alter the basic physics of the
980: collisionless shock, if a mean field is present. Milosavljevic and
981: Nakar (2006b) argue accelerated particles streaming into the {\it
982: upstream} magnetized medium can drive long wavelength, magnetized
983: turbulence with $\delta B /B \gg1$, which might persist into the
984: downstream and provide the magnetization required in phenomenological
985: models of GRB and PWN emission. Then the shock mediated by Weibel
986: turbulence becomes a subshock within a much larger extended structure,
987: responsible only for thermalizing the bulk of the flow and injecting
988: the particles that are Fermi accelerated in the turbulence generated
989: by high energy particle streaming. This is a relativistic version of
990: Bell's (1978) (also see Bell 2004, 2005) picture of particle
991: acceleration in non-relativistic shocks.
992:
993: Finally, we briefly mention the extension of this theory to
994: electron-ion plasmas. Let us presume initially that the ions and
995: electrons are isotropic but remain decoupled, i.e., a two-temperature
996: relativistic plasma. Equation (\ref{eq:chi_limit}) becomes
997: \begin{equation}\label{eq:i-e chi_limit}
998: \chi \approx i\frac {\pi} 4 \frac {\omega_{p,i}^2 + \omega_{p,e}^2}{|k|c\omega},
999: \end{equation}
1000: where $\omega_{p, i}^2 = 4\pi n_i e^2/\gamma_i m_i$ is the plasma
1001: frequency of the ions, $\gamma_i$ is the Lorentz factor associated
1002: with the ion temperature, $\omega_{p, e}^2 = 4\pi n_e e^2/\gamma_e
1003: m_e$ is the plasma frequency of the electrons, and $\gamma_e$ is the
1004: Lorentz factor associated with the electron temperature. Depending on
1005: the relative values of the electron temperature and ion temperature,
1006: one term may dominate. However, initial large-scale simulations of
1007: ion-electron collisionless shocks suggest that both reach roughly
1008: equipartition with each other (Spitkovsky 2007), thereby reproducing
1009: the physics of the $e^{\pm}$ shock. In this case, the relativistic
1010: electrons and ions contribute equally to the decay rate because
1011: thermal equipartition prevails, i.e., $m_e \gamma_e = m_i
1012: \gamma_i$. Thus, the electron-ion plasma has the same dynamics as the
1013: electron-positron plasma.
1014:
1015: \acknowledgements
1016:
1017: We thank S. Cowley, D. Kocelski, M. Milosalavjic, A. Pe'er and E.
1018: Quataert for useful discussions. P.C. and J.A. thank the Institute for
1019: Advanced Study for its hospitality; P.C. also thanks the Canadian
1020: Institute for Theoretical Astrophysics for similar hospitality. P.C.
1021: is supported by the Miller Institute for Basic Research. J.A. has
1022: benefited from the support of NSF grant AST-0507813, NASA grant
1023: NNG06GI08G, and DOE grant DE-FC02-06ER41453, all at UC Berkeley; by
1024: the Department of Energy contract to the Stanford Linear Accelerator
1025: Center no. DE-AC3-76SF00515; and by the taxpayers of California. A.S. is pleased to acknowledge that the simulations reported on
1026: in this paper were substantially performed at the TIGRESS high
1027: performance computer
1028: center at Princeton University which is jointly supported by the Princeton
1029: Institute for Computational Science and Engineering and the Princeton
1030: University Office of
1031: Information Technology.
1032: \appendix
1033: \section{Susceptibility in Two and Three Dimensions}\label{sec:susceptibility}
1034:
1035: In this section, we solve the susceptibility (eq.[\ref{eq:susceptibility}]) for two and three dimensional plasma. The 3D case
1036: (\S\ref{sec:3dresult}) has been previously solved by Mikhailovskii
1037: (1979) and we reproduce his result here for completeness. In
1038: addition, we present the 2D case (\S\ref{sec:2dresult}), which
1039: allows for a comparison to the two-dimensional simulations reported in
1040: this paper.
1041:
1042: \subsection{Three Dimensions}\label{sec:3dresult}
1043:
1044: After summing over the electrons and positrons, equation (\ref{eq:susceptibility}) is
1045: \begin{equation}\label{eq:susceptibility 3d}
1046: 4 \pi \chi = \frac {\omega_{p,{\rm NR}}^2}{\omega^2}
1047: \int v_x \left(\frac {d} {dp_x} + \frac {kv_x}{\omega - kv_y}
1048: \frac {d}{dp_y}\right)\frac {f_0}{n} d^3p,
1049: \end{equation}
1050: where $n$ is the number density of electrons and positions. Since
1051: $f_0$ is isotropic and therefore independent of angle, we orient the
1052: spherical integral in a {\it non-standard manner} so that the pole
1053: points along the k-vector, i.e., the y-axis. We find
1054: $\boldsymbol{\hat{p}}\cdot \boldsymbol{\hat{y}} = \cos\theta$ and
1055: $\boldsymbol{\hat{p}}\cdot \boldsymbol{\hat{x}} = \sin\theta\sin\phi$.
1056: So equation (\ref{eq:susceptibility 3d}) becomes
1057: \begin{equation}
1058: 4 \pi \chi = \frac {\omega_{p,{\rm NR}}^2}{\omega^2n}
1059: \int v\sin^2\theta\sin^2\phi \left(1 + \frac {\cos\theta}{\omega/kv - \cos\theta}\right)\frac {df_0} {dp} p^2dpd\Omega,
1060: \end{equation}
1061: where $d\Omega = \sin\theta d\theta d\phi$. Performing the integral
1062: over $\phi$ and making the substitution $\zeta = \cos\theta$, we find
1063: \begin{equation}
1064: 4 \pi \chi = \frac {\pi\omega_{p,{\rm NR}}^2}{\omega^2n}
1065: \int v\left(1-\zeta^2\right) \left(1 + \frac {\zeta}{\omega/kv - \zeta}\right)\frac {df_0} {dp} p^2dpd\zeta.
1066: \end{equation}
1067: Integrating over $\zeta$ from -1 to 1, we find
1068: \begin{eqnarray}
1069: 4 \pi \chi & = & \frac {2\pi\omega_{p,{\rm NR}}^2}{\omega^2n} \int p^2dp \frac {df_0}{dp} v{\Bigg \{} \left(\frac
1070: {\omega} {kv}\right)^2 - \frac{\omega}{kv}\left[1 - \left(\frac{\omega} {kv}\right)^2\right]
1071: \nonumber \\
1072: & \hspace*{0cm} & \hspace{1cm} \times \left[ \frac{1}{ 2} \log \left(-\frac
1073: {1 + \omega/kv}{1 - \omega/kv}\right)\right] {\Bigg \}}.
1074: \label{eq:worked out suscept1}
1075: \end{eqnarray}
1076: Note that the logarithmic function will give an imaginary part when
1077: $\omega_r/kc\beta < 1$. This implies that waves whose phase velocity,
1078: $\omega_r/k$, is small compared to the thermal speed of the background
1079: particles, $c\beta$, will be damped (also see \S\ref{sec:theory}). We
1080: pull this imaginary component out of the equation, which makes this damping
1081: more explicit, to find:
1082: \begin{eqnarray}
1083: 4 \pi \chi & = & \frac {2\pi\omega_{p,{\rm NR}}^2}{\omega^2n} \int p^2dp \frac {df_0}{dp} v{\Bigg \{} \left(\frac
1084: {\omega} {kv}\right)^2 - \frac{\omega}{kv}\left[1 - \left(\frac{\omega} {kv}\right)^2\right]
1085: \nonumber \\
1086: & \hspace*{0cm} & \hspace{1cm} \times \left[ \frac{1}{ 4} \log \left(\frac
1087: {1 + \omega/kv}{1 - \omega/kv}\right)^2 -
1088: i\frac {\pi}{2}
1089: \Theta\left(k^2 v^2 - \omega_r^2 \right)\right] {\Bigg \}}
1090: , \label{eq:worked out suscept}
1091: \end{eqnarray}
1092: where $\Theta$ is the unit step function (i.e., $\Theta(z) = 1$ for
1093: $\Re(z) \geq 0$ and $\Theta(z) = 0$ for $\Re(z) < 0$).
1094: Equation (\ref{eq:worked out suscept}) is precisely Mikhailovskii
1095: (1979)'s result.
1096: We now apply a three dimensional relativistic Maxwellian $f_0 \propto
1097: \exp\left(-{E}/kT\right)$, which is appropriately normalized, $\int
1098: d^3p f_0 = n$, and solve equation (\ref{eq:worked out suscept})
1099: numerically. For the ultrarelativistic case $v\approx c$, we find a
1100: simple form in the limit $\omega_r \ll kc$:
1101: \begin{equation}\label{eq:3dchi}
1102: 4\pi\chi \approx i\frac {\pi} 4 \frac {\omega_{p}^2}{|k|c\omega}.
1103: \end{equation}
1104:
1105: \subsection{Two Dimensions}\label{sec:2dresult}
1106:
1107: Starting from equation (\ref{eq:susceptibility 3d}), we assume $f_0 =
1108: f(p_{2d})g(p_z)$, where $p_{2d} = \sqrt{p_x^2 + p_y^2}$ and perform
1109: the integral over $p_z$. The resulting two dimensional analogue of
1110: equation (\ref{eq:susceptibility 3d}) is
1111: \begin{equation}\label{eq:susceptibility 2d}
1112: 4 \pi \chi = \frac {\omega_{p,{\rm NR}}^2}{\omega^2}
1113: \int v_x \left(\frac {d} {dp_x} + \frac {kv_x}{\omega - kv_y}
1114: \frac {d}{dp_y}\right)\frac {f} n d^2p,
1115: \end{equation}
1116: where we have dropped the subscript ``$2d$'' from $p$. Defining $p_x =
1117: p\cos\theta$, $p_y = p\sin\theta$, and similarly for $v_x$ and $v_y$,
1118: we find:
1119: \begin{equation}
1120: 4\pi\chi= \frac {\omega_{p,{\rm NR}}^2}{\omega^2 n}
1121: \int pdp \int_0^{2\pi} d\theta v\cos^2\theta \left(1 + \frac
1122: {\sin\theta}{(\omega/kv) - \sin\theta} \right)\frac{df}{dp}.
1123: \end{equation}
1124: We may transform the $\theta$ integral from $0$ to $2\pi$ to a contour
1125: integral over the unit circle by making the appropriate substitutions
1126: (Carrier, Krook, \& Pearson 1983)
1127: \begin{eqnarray}
1128: \cos\theta &\rightarrow& \frac 1 2\left(z + z^{-1}\right),
1129: \sin\theta \rightarrow \frac 1 {2i}\left(z - z^{-1}\right) \\
1130: d\theta &\rightarrow& \frac {dz} {iz},
1131: \int_0^{2\pi} \rightarrow \int_{\Gamma},
1132: \end{eqnarray}
1133: where $\Gamma$ is the unit circle. After a bit of algebra, we find
1134: \begin{equation}
1135: 4\pi\chi= \frac {\omega_{p,{\rm NR}}^2}{\omega^2n}
1136: \int v \frac{df}{dp} pdp \int_{\Gamma} \frac {-i} 4 dz
1137: \left(z^2 + 2 + z^{-2}\right)\frac
1138: {(2i\omega/kv)}{(2i\omega/kv)z - z^2 + 1}.
1139: \end{equation}
1140: The singular points in this equation are $z_{\pm} = (i\omega/kv) \mp
1141: \sqrt{1 -(\omega/kv)^2}$. We perform the contour integral by noting
1142: that only the $z_-$ root contributes for $(\omega/kv)^2 < 1$:
1143: \begin{equation}\label{eq:2-d susceptibility}
1144: 4\pi\chi = -i\frac {2\pi\omega_{p,{\rm NR}}^2}{\omega k n}
1145: \int \sqrt{1 - \left(\frac{\omega}{kv}\right)^2} \frac{df}{dp} pdp.
1146: \end{equation}
1147: We apply a two dimensional relativistic Maxwellian $f \propto
1148: \exp\left(-{E}/kT\right)$, where $f$ is appropriately normalized, i.e.
1149: $\int f d^2p = n$. Expanding to lowest order in $\omega/kv$, we find:
1150: \begin{equation}\label{eq:2-d result}
1151: 4\pi\chi \approx i\frac {\omega_{p}^2}{|k|c\omega}.
1152: \end{equation}
1153:
1154:
1155: \section{Decay of an Initial Field of Fluctuations}\label{sec:fluctuation dissipation theorem}
1156:
1157: The plasma susceptibilities (eq.[\ref{eq:chi_limit}]) define the
1158: linear response of the plasma. These susceptibilities are complex and
1159: hence the electric permittivity $\epsilon = 1 + 4\pi\chi$ is also
1160: complex. The implications of Poynting's theorem for the propagation
1161: of small amplitude waves and fluctuations in such a medium have been
1162: described in many texts and monographs (e.g., Bekefi 1966, Melrose and
1163: McPhedran 1991, Stix 1992). For completeness, we give a brief discussion
1164: of the theory behind expression (\ref{eq:damping}).
1165:
1166: We begin with Poynting's theorem:
1167: \begin{equation}
1168: \frac{\partial}{\partial t} \left(\frac{\delta B^2 + \delta E^2}{8\pi}
1169: \right) + {\boldsymbol \nabla}\cdot{\boldsymbol S} = - {\delta
1170: \boldsymbol j} \cdot \delta {\boldsymbol E},
1171: \end{equation}
1172: where ${\boldsymbol S} = (c/4\pi){\boldsymbol E}\times{\boldsymbol B}$
1173: is the Poynting vector. We assume field energy is distributed
1174: uniformly in our uniform plasma, so ${\boldsymbol
1175: \nabla}\cdot{\boldsymbol S} = 0$, i.e., there is no transport of
1176: energy spatially. In addition, for nonpropagating Weibel modes, $\delta
1177: E \ll \delta B$. Thus, we find a simplified expression:
1178: \begin{equation}
1179: \frac{\partial}{\partial t} \left(\frac{\delta B^2}{8\pi} \right) = -
1180: {\delta \boldsymbol j} \cdot \delta {\boldsymbol E}.
1181: \end{equation}
1182: Writing the fields with truncated amplitudes
1183: \begin{equation} \delta B_{TV} ( {\boldsymbol r}, t) = \left\{
1184: \begin{array}{rl}
1185: \delta B, \; t \in (-T/2, T/2), \; |{\boldsymbol r}| \in V, \\
1186: 0, \; t \not\in (-T/2, T/2), \; |{\boldsymbol r}| \not\in V,
1187: \end{array}
1188: \right.
1189: \end{equation}
1190: with $\delta B_{TV} = 0$ when the coordinates are outside of the
1191: volume, $V$, and time is outside of the interval $(-T/2, T/2)$. Thus,
1192: we can define space-time averages of the fields while still expressing
1193: them in terms of convergent Fourier transforms. We write the Fourier
1194: transforms as
1195: \begin{equation}
1196: \delta j_{k\omega} = \int d^3k \int_{-\infty}^{\infty}d\omega\,\delta j_{TV}
1197: \exp\left(i{\boldsymbol k}\cdot{\boldsymbol r} - i\omega t\right).
1198: \end{equation}
1199: We express the field energy density and the $\delta j \cdot \delta E$
1200: work in terms of the Fourier amplitudes and then average over time and
1201: space. Taking $T$ and $V$ to $\infty$, we integrate over the resulting
1202: $\delta$-functions to find
1203: \begin{equation}
1204: \frac{\partial}{\partial t} \left \langle \frac{\delta B^2}{8\pi} \right \rangle =
1205: \int d^3k\int d\omega\frac 1 2
1206: \langle \delta j_{k\omega } \delta E_{k\omega}^* + c.c. \rangle,
1207: \end{equation}
1208: where $\langle \rangle$ represents the space time average over the
1209: fluctuation wavelengths and variability times and $c.c.$ is the
1210: complex conjugate. We apply the same Fourier transform to $\delta B$
1211: and apply the same average over time and space. We use the linear
1212: response (eq.[\ref{eq:j2E}]) $\delta E_{k\omega} = (i / \chi \omega)
1213: \delta j_{k\omega}$ to find
1214: \begin{equation}
1215: \frac 1 {8\pi}\frac{\partial \langle |\delta B_{k\omega}|^2 \rangle}{\partial t}
1216: = -\Im\left(\omega \chi\right)^{-1} \langle |\delta j_{k\omega} |^2 \rangle.
1217: \end{equation}
1218: Now using $\delta j_{k\omega} = ikc \delta B_{k\omega}/4\pi$,
1219: we find
1220: \begin{equation}\label{eq:app_damping}
1221: \frac{\partial \langle |\delta B_{k\omega}|^2 \rangle }{\partial t}
1222: = -2 \gamma_{k\omega} \langle |\delta B_{k\omega} |^2 \rangle,
1223: \end{equation}
1224: where $\gamma_{k\omega}$ is the damping rate or
1225: \begin{equation}\label{eq:app_general gamma}
1226: \gamma_{k\omega} = \frac {\left(kc\right)^2}{\omega} \Im\left( 4\pi\chi\right)^{-1}.
1227: \end{equation}
1228: Equation (\ref{eq:app_damping}) is a specific form of the
1229: fluctuation-dissipation theorem (Thompson and Hubbard 1960) from which
1230: the same result can be derived.
1231:
1232: Equations (\ref{eq:app_damping}) and (\ref{eq:app_general gamma})
1233: define the evolution of magnetic energy in terms of the linear
1234: response. For notational simplicity we reduce $k\omega \rightarrow k$
1235: in the rest of the text. We obtain simple forms for $\gamma_k$, using
1236: the asymptotic forms of $4\pi\chi$ from equation (\ref{eq:chi_limit}) for
1237: 3D and 2D. We find for $kc \ll \omega_p$:
1238: \begin{equation}\label{eq:app_gamma}
1239: \gamma_{k} = \left\{\begin{array}{ll} \frac {|kc|^3} {\omega_{p}^2} & \textrm{2D}
1240: \\ \frac {4} {\pi} \frac{|kc|^3} {\omega_{p}^2}& \textrm{3D}
1241: \end{array}\right. .
1242: \end{equation}
1243: The cubic dependence on wavenumber in equation (\ref{eq:app_gamma})
1244: suggests short wavelength modes are very strongly damped, while longer
1245: wavelength modes can survive much longer.
1246: %In addition, the 2D and 3D
1247: %damping rates are the same up to a overall multiplicative constant of
1248: %order unity.
1249:
1250: Taking the general form of $4\pi\chi$ for 2D and 3D from equation
1251: (\ref{eq:worked out suscept}) and (\ref{eq:2-d susceptibility}), we
1252: numerically compute the damping rate from equation
1253: (\ref{eq:app_general gamma}) for Weibel modes where $\omega_r = 0$, because of the
1254: non-propagating nature of the downstream magnetic clumps. We
1255: show the results in Figure \ref{fig:damping rates}. Also
1256: plotted are the simple forms for $\gamma_k$ from equation
1257: (\ref{eq:app_gamma}). Though, the asymptotic forms are only valid for
1258: $kc/\omega_{p} \ll 1$, the numerical and asymptotic results are in
1259: excellent agreement throughout. Thus for simplicity, we use equation
1260: (\ref{eq:app_gamma}) (also eq.[\ref{eq:gamma}] in the main body) for
1261: the damping rates.
1262:
1263: \begin{figure}[H]
1264: \plotone{f6.eps}
1265: \caption{Damping rates of magnetic field as a function of $kc/\omega_{p}$. We plot
1266: the damping rates numerically computed from equation
1267: (\ref{eq:worked out suscept}) and (\ref{eq:2-d susceptibility}) for 3D
1268: (thick solid line) and 2D (thin solid line) respectively. Also
1269: overplotted are the analytic forms of these expressions from
1270: equation (\ref{eq:gamma}) for 2D (thin dashed line) and 3D (thick
1271: dashed line).}
1272: \label{fig:damping rates}
1273: \end{figure}
1274:
1275: %We find excellent agreement between the numerical evaluation of
1276: %$\gamma_k$ for $\omega_r = 0$ and equation (\ref{eq:app_gamma}) (see
1277: %Appendix A), so we use equation (\ref{eq:app_gamma}) in the rest of this
1278: %paper.
1279:
1280: \begin{references}
1281:
1282: \reference{}
1283: Arons, J., Norman, C. A., \& Max, C. E. 1977, Phys. Fluids, 20, 1302
1284:
1285: %\reference{}
1286: %Barnes, A. 1968, \apj, 154, 751
1287:
1288: \reference{}
1289: Bekefi, G. 1966, ``Radiation Processes in Plasmas'' (John Wiley: New York)
1290:
1291: \reference{}
1292: Bell, A. R. 1978, \mnras, 182, 147
1293:
1294: \reference{}
1295: Bell, A. R. 2004, \mnras, 353, 550
1296:
1297: \reference{}
1298: Bell, A. R. 2005, \mnras, 358, 181
1299:
1300: \reference{}
1301: Birdsall, C. K. \& Langdon, A. B. 1991, ``Plasma Physics via Computer Simulations'' (McGraw-Hill: New York)
1302:
1303: \reference{}
1304: Buneman, O. 1993 in ``Computer Space Plasma Physics'', Terra Scientific, Tokyo, 67
1305:
1306: \reference{}
1307: Carrier, G. F., Krook, M., \& Pearson, C. E. 1983, ``Functions of a Complex Variable : Theory and Technique'' (Hod Books: Ithaca)
1308:
1309: \reference{}
1310: Coroniti, F.~V. 1990, \apj, 349, 538
1311:
1312: \reference{}
1313: Davidson, R. C., Hammer, D. A., Haber, I., \& Wagner, C. E. 1972, Phys. Fluids, 15, 317
1314:
1315: \reference{}
1316: Frederiksen, J. T., Hededal, C. B., Haugb\`{o}lle, T., \& Nordlund, �Å.
1317: 2004, \apj, 608, L13
1318:
1319: \reference{}
1320: Gallant, Y. A., Hoshino, M., Langdon, A. B., Arons, J., \& Max, C. E. 1992, \apj, 391, 73
1321:
1322: \reference{}
1323: Gruzinov, A. 2001a, submitted to \apj, astro-ph/0111321
1324:
1325: \reference{}
1326: Gruzinov, A. 2001b, \apj, 563, L15
1327:
1328: \reference{}
1329: Gruzinov, A. \& Waxman E. 1999, \apj, 511, 852
1330:
1331: \reference{}
1332: Hammett, G. W., Dorland, W., \& Perkins, F. W. 1992, Phys. Fluids B, 4, 2052
1333:
1334: \reference{}
1335: Hededal, C. B., Trier Frederiksen, J., Haugboelle, T., \& Nordlund,
1336: A. 2005, Neutrinos and Explosive Events in the Universe, Proceedings of
1337: the 14th Course of the International School of Cosmic Rays
1338: Astrophysics, a NATO Advanced Study Institute, held in Erice, Italy,
1339: 2-13 July 2004. Edited by Maurice M. Shapiro, Stanev Todor, and John
1340: P. Wefel.
1341:
1342: \reference{}
1343: Jackson, J.~D. 1999, ``Classical Electrodynamics'' (John Wiley \& Sons: New York)
1344:
1345: \reference{}
1346: Kato, T. N. 2005, Phys. Plasmas, 12, 80705
1347:
1348: \reference{}
1349: Katz, B., Keshet, U., \& Waxman, E. 2007, \apj, 655, 375
1350:
1351: \reference{}
1352: Medvedev, M.~V. \& Loeb, A. 1999, \apj, 526, 697
1353:
1354: \reference{}
1355: Medvedev, M. V., Fiore, M., Fonseca, R. A., Silva,
1356: L. O.; Mori, W. B. 2005, \apj, 618, L75
1357:
1358: \reference{}
1359: Melrose, D.B., and McPhedran, R.C. 1991, ``Electromagnetic Processes in
1360: Dispersive Media'' (Cambridge: Cambridge University Press)
1361:
1362: \reference{}
1363: Milosavljevic, M., Nakar, E., \& Spitkovsky, A. 2006, \apj, 637, 765
1364:
1365: \reference{}
1366: Milosavljevic, M. \& Nakar, E. 2006a, \apj, 641, 978
1367:
1368: \reference{}
1369: Milosavljevic, M. \& Nakar, E. 2006b, \apj, 651, 979
1370:
1371: %\reference{}
1372: %Moiseev, S.~S., \& Sagdeev, R.~Z. 1963, J. Nucl. Energy C, 5, 43
1373:
1374: \reference{}
1375: Mikhailovskii, A. B. 1979, Plasma Phys., 22, 133
1376:
1377: \reference{}
1378: Nishikawa, K.-I., Hardee, P., Richardson, G., Preece, R., Sol, H., \&
1379: Fishman, G. J. 2003, \apj, 595, 555
1380:
1381: \reference{}
1382: Nishikawa, K.-I., Hardee, P., Richardson, G., Preece, R., Sol, H., \&
1383: Fishman, G. J. 2005, \apj, 622, 927
1384:
1385: \reference{}
1386: Pe'er, A. \& Zhang, B. 2006, \apj, 653, 454
1387:
1388: \reference{}
1389: Piran, T. 2005a, Rev. of Modern Phys., 76, 1143
1390:
1391: \reference{}
1392: Piran, T. 2005b, Magnetic Fields in the Universe, Angra dos Reis,
1393: Brazil, Nov. 29-Dec 3, 2004, Ed. E. de Gouveia del Pino, AIP
1394: Conference Proceedings, v784 (New York:AIP), p164
1395:
1396: \reference{}
1397: Rossi, E. \& Rees, M. J. 2003, \mnras, 339, 881
1398:
1399: \reference{}
1400: Silva, L. O., Fonseca, R. A., Tonge, J. W., Dawson, J. M., Mori,
1401: W. B., \& Medvedev, M. V. 2003, \apj, 596, 121
1402:
1403: \reference{}
1404: Spitkovsky, A. 2005, AIP Conf. Proc, 801, 345; astro-ph/0603211
1405:
1406: \reference{}
1407: Spitkovsky, A. 2007, submitted to \apjl; arXiv:0706.3126
1408:
1409: \reference{}
1410: Stix, T.H. 1992, ``Waves in Plasmas'' (American Institute of Physics: New York)
1411:
1412: \reference{}
1413: Thompson, W. B. \& Hubbard, J. 1960, Rev. of Modern Phys., 32, 714
1414:
1415: \reference{}
1416: Weibel, E.~S. 1959, \prl, 2, 83
1417:
1418: %\reference{}
1419: %Wright, T. P. \& Hadley, G. R. 1975, \pra, 12, 686
1420:
1421: \reference{}
1422: Yoon, P.~H., \& Davidson, R.~C. 1987, \pra, 35, 2718
1423:
1424: \end{references}
1425:
1426: %\clearpage
1427:
1428:
1429:
1430:
1431: \end{document}