1: %% The command below calls the preprint style
2: %% which will produce a one-column, single-spaced document.
3:
4: %%\documentclass[12pt,preprint]{aastex}
5: %% manuscript produces a one-column, double-spaced document:
6: %%\documentclass[12pt,manuscript]{aastex}
7: \documentclass[apj]{emulateapj}
8: %%\documentclass[apjl]{emulateapj}
9: %%\documentclass[12pt,preprint2]{aastex}
10:
11: %% preprint2 produces a double-column, single-spaced document:
12: %%\documentclass[preprint2]{aastex}
13:
14: %% Sometimes a paper's abstract is too long to fit on the
15: %% title page in preprint2 mode. When that is the case,
16: %% use the longabstract style option.
17: %% \documentclass[preprint2,longabstract]{aastex}
18:
19: %%\usepackage{ulem}
20: %%\usepackage{color}
21:
22: \newcount\listnorom
23: \listnorom=0
24: \newcommand\listromanDE{\global\advance \listnorom by 1
25: {\lowercase\expandafter{(\romannumeral\listnorom)}\ }}
26: \newcommand\newlistroman{\listnorom=0}
27: %
28:
29: \newcommand{\Tgp}{\tau_\mathrm{gp}}
30: \newcommand{\Eth}{E_\mathrm{th}}
31: \newcommand{\xx}[1]{\!\times\!10^{#1}}
32: \newcommand{\red}{\textcolor{red}}
33: \newcommand{\blue}{\textcolor{blue}}
34: \newcommand{\green}{\textcolor{green}}
35: \newcommand{\FP}{Analytic Precursor Approximation}
36: \newcommand{\FPshort}{APA}
37:
38: \newcommand{\Lcell}{L_\mathrm{cell}}
39: \newcommand{\tstep}{t_\mathrm{step}}
40: \newcommand{\Vsk}{v_\mathrm{sk}}
41: \newcommand{\IonPlasmafreq}{\omega_{pi}}
42: \newcommand{\ElPlasmafreq}{\omega_{pe}}
43: \newcommand{\plasmafreq}{\omega_{pi}}
44: \newcommand{\protongyrofreq}{\omega_{\mathrm{cp}}}
45: \newcommand{\MFA}{magnetic field amplification}
46: \newcommand{\rel}{relativistic}
47: \newcommand{\NL}{nonlinear}
48: \newcommand{\CR}{cosmic ray}
49: \newcommand{\CRs}{cosmic rays}
50: \newcommand{\nonrel}{non\-rel\-a\-tiv\-is\-tic}
51: \newcommand{\transrel}{trans-rel\-a\-tiv\-is\-tic}
52: \newcommand{\ultrarel}{ul\-tra-rel\-a\-tiv\-is\-tic}
53: \newcommand{\mc}{Monte Carlo}
54: \newcommand{\MC}{Monte Carlo}
55: \newcommand{\syn}{synchrotron}
56: \newcommand{\synch}{synchrotron}
57: \newcommand{\pion}{pion-decay}
58: \newcommand{\IC}{inverse-Compton}
59: \newcommand{\kmps}{km s$^{-1}$}
60: \newcommand{\pcc}{cm$^{-3}$}
61: \newcommand{\muG}{$\mu$G}
62:
63: \newcommand{\SC}{self-consistent}
64: \newcommand{\SCly}{self-consistently}
65: \newcommand{\Pcr}{P_\mathrm{cr}}
66: \newcommand{\Pcrtwo}{P_\mathrm{cr2}}
67: \newcommand{\Pcrhat}{P_\mathrm{cr}}
68: \newcommand{\Pth}{P_\mathrm{th}}
69: \newcommand{\Pthzero}{P_\mathrm{th0}}
70: \newcommand{\Pwzero}{P_\mathrm{w0}}
71: \newcommand{\Pwone}{P_\mathrm{w1}}
72: \newcommand{\Pwtwo}{P_\mathrm{w2}}
73: \newcommand{\Fwzero}{F_\mathrm{w0}}
74: \newcommand{\Fwone}{F_\mathrm{w1}}
75: \newcommand{\Fwtwo}{F_\mathrm{w2}}
76: \newcommand{\Ppzero}{P_\mathrm{p0}}
77: \newcommand{\Ppone}{P_\mathrm{p1}}
78: \newcommand{\Pptwo}{P_\mathrm{p2}}
79: \newcommand{\wpzero}{w_\mathrm{p0}}
80: \newcommand{\wpone}{w_\mathrm{p1}}
81: \newcommand{\wptwo}{w_\mathrm{p2}}
82: \newcommand{\Rtot}{r_\mathrm{tot}}
83: \newcommand{\rtot}{r_\mathrm{tot}}
84: \newcommand{\Rsub}{r_\mathrm{sub}}
85: \newcommand{\rsub}{r_\mathrm{sub}}
86: \newcommand{\Mazero}{M_\mathrm{A0}}
87: \newcommand{\Mszero}{M_\mathrm{s0}}
88: \newcommand{\Msone}{M_\mathrm{s1}}
89: \newcommand{\vth}{v_\mathrm{th}}
90: \newcommand{\vwt}{v_\mathrm{wt}}
91: \newcommand{\vazero}{v_\mathrm{A0}}
92: \newcommand{\xFEB}{x_\mathrm{FEB}}
93: \newcommand{\xFP}{x_\mathrm{APA}}
94: \newcommand{\xtr}{x_\mathrm{tr}}
95: \newcommand{\fcr}{f_\mathrm{cr}}
96: \newcommand{\AccEff}{{\cal E}_\mathrm{cr}}
97: \newcommand{\xicr}{\xi_\mathrm{cr}}
98: \newcommand{\pmax}{p_\mathrm{max}}
99: \newcommand{\Emax}{E_\mathrm{max}}
100: \newcommand{\rgTeV}{r_\mathrm{g}(\mathrm{TeV})}
101: \newcommand{\rgzero}{r_\mathrm{g0}}
102: \newcommand{\tacc}{\tau_\mathrm{acc}}
103: \newcommand{\fAlf}{f_\mathrm{Alf}}
104: \newcommand{\Beff}{B_\mathrm{eff}}
105: \newcommand{\Bism}{B_\mathrm{eff}}
106: \newcommand{\Befftwo}{B_\mathrm{eff2}}
107: \newcommand{\Btrend}{B_\mathrm{trend}}
108: \newcommand{\Qesc}{Q_\mathrm{esc}}
109: \newcommand{\qesc}{q_\mathrm{esc}}
110: \newcommand{\taucoll}{\tau_\mathrm{coll}}
111: \newcommand{\taucomp}{\tau_\mathrm{comp}}
112: \newcommand{\momentumflux}{\Phi_P}
113: \newcommand{\upstreammomentumflux}{\Phi_\mathrm{P0}}
114: \newcommand{\downstreammomentumflux}{\Phi_\mathrm{P2}}
115: \newcommand{\energyflux}{\Phi_E}
116: \newcommand{\upstreamenergyflux}{\Phi_\mathrm{E0}}
117: \newcommand{\downstreamenergyflux}{\Phi_\mathrm{E2}}
118: \newcommand{\gammabar}{\bar{\gamma}}
119: \newcommand{\deltabar}{\bar{\delta}}
120: \newcommand{\pvector}{{\bf p}}
121: \newcommand{\vvector}{{\bf v}}
122:
123: \newcommand{\coldism}{cold ISM ($T_0=10^4$~K)}
124: \newcommand{\hotism}{hot ISM ($T_0=10^6$~K)}
125: \newcommand{\coldismNoT}{cold ISM}
126: \newcommand{\hotismNoT}{hot ISM}
127:
128:
129: \newcommand{\avemail}{avladim@ncsu.edu}
130: \newcommand{\abemail}{byk@astro.ioffe.ru}
131: \newcommand{\deemail}{don\_ellison@ncsu.edu}
132: \newcommand{\heatpar}{\alpha_H}
133: \newcommand{\Alf}{Alfv\'{e}n}
134: \newcommand{\Lbar}{L}
135:
136: \shorttitle{Turbulence Dissipation in DSA}
137: \shortauthors{Vladimirov, Bykov, \& Ellison}
138:
139: \begin{document}
140:
141: \title{Turbulence Dissipation and Particle Injection in Non-Linear
142: Diffusive Shock Acceleration with Magnetic Field Amplification}
143:
144: \author{Andrey E. Vladimirov} \affil{Physics Department, North Carolina
145: State University, Box 8202, Raleigh, NC 27695; \avemail}
146:
147: \author{Andrei M. Bykov} \affil{Department of Theoretical Astrophysics,
148: Ioffe Physical-Technical Institute, St. Petersburg, Russia; \abemail}
149:
150: \and
151:
152: \author{Donald C. Ellison} \affil{Physics Department, North Carolina
153: State University, Box 8202, Raleigh, NC 27695; \deemail}
154:
155:
156: \begin{abstract}
157:
158: The highly amplified magnetic fields suggested by observations of some
159: supernova remnant (SNR) shells are most likely an intrinsic part of
160: efficient particle acceleration by shocks.
161: %
162: This strong turbulence, which may result from cosmic ray driven instabilities,
163: both resonant and non-resonant, in the shock precursor, is certain to
164: play a critical role in \SC, \NL\ models of strong, \CR\ modified shocks.
165: %
166: Here we present a Monte Carlo model of nonlinear diffusive shock
167: acceleration (DSA) accounting for magnetic field amplification through
168: resonant instabilities induced by accelerated particles, and including
169: the effects of dissipation of
170: turbulence upstream of a shock and the subsequent precursor plasma
171: heating. Feedback effects between the plasma heating due to
172: turbulence dissipation and particle injection are strong, adding to the
173: \NL\ nature of efficient DSA.
174: %
175: Describing the turbulence damping in a parameterized way, we reach two
176: important results: first, for conditions
177: typical of supernova remnant shocks, even a small
178: amount of dissipated turbulence energy ($\sim 10\%$) is sufficient to
179: significantly heat the precursor plasma, and second,
180: the heating upstream of the shock leads to an increase in the
181: injection of thermal particles at the subshock by a factor
182: of several.
183: In our results, the response of the non-linear shock
184: structure to the boost in particle injection prevented
185: the efficiency of particle acceleration and magnetic
186: field amplification from increasing.
187: We argue, however, that more advanced (possibly, non-resonant) models
188: of turbulence generation and dissipation may lead to
189: a scenario in which particle injection boost due
190: to turbulence dissipation results in more efficient
191: acceleration and even stronger amplified magnetic fields than
192: without the dissipation.
193:
194: \end{abstract}
195:
196: \keywords{acceleration of particles -- cosmic rays -- supernova remnants
197: -- magnetic fields -- turbulence -- shock waves}
198:
199:
200: \section{Introduction}
201: Observations of young supernova remnants (SNRs) suggest that strong
202: collisionless shocks can simultaneously
203: %
204: place a large fraction of the shock ram kinetic energy into \rel\ protons
205: %
206: \citep[e.g.,][]{BE87,MD2001, WarrenEtal2005}
207: %
208: and amplify the ambient turbulent magnetic field by large factors
209: %
210: \citep[e.g.,][]{Cowsik80, RE92, BambaEtal2003, BKV2003, VL2003,
211: UA2008}.
212: %
213: This coupling of diffusive shock acceleration (DSA) and magnetic field
214: amplification (MFA) is critically important because the self-generated
215: magnetic field largely determines the efficiency of DSA, the maximum
216: particle energy a given shock can produce, and the \syn\ emission from
217: radiating electrons.
218:
219: The generation and dissipation of strong MHD turbulence in
220: collisionless, multi-fluid plasmas is a complex process.
221: %
222: Different \NL\ approaches to the modeling of the large scale
223: structure of a shock undergoing efficient cosmic ray acceleration
224: \citep[e.g.,][]{MV82,AB86,BL2001,AB2006,VEB2006,PLM2006,ZPV2008} have
225: predicted the presence of strong MHD turbulence in the shock
226: precursor.
227: %
228: However, an exact modeling of the shock structure in a turbulent medium,
229: including nonthermal particle injection and acceleration, requires a
230: nonperturbative, self-consistent description of a multi-component and
231: multi-scale system including the strong MHD-turbulence dynamics.
232: %
233: While a number of analytic models describing resonant and non-resonant
234: amplification and damping of magnetic fluctuations have been proposed,
235: these generally rely on the quasi-linear approximation that the
236: fluctuations are small compared to the background magnetic field, i.e.,
237: $\Delta B \ll B_0$ \citep[e.g.,][]{Wentzel74,BT2005,Kulsrud2005}.
238: %
239: No consistent analytic description of magnetic turbulence generation
240: with $\Delta B \gtrsim B_0$ exists. For these reasons, numerical models
241: with varying ranges of applicability have been proposed which offer a
242: compromise between completeness and speed
243: \citep[e.g.,][]{Bell2004,AB2006,VEB2006,ZPV2008}.
244:
245: In principle, the problem can be solved completely with few assumptions
246: and approximations with particle-in-cell (PIC) simulations
247: \citep[e.g.,][]{Bell2004,Spitkovsky2008,NPS2008}
248: or, in the assumption that electrons are not dynamically
249: important, by hybrid models \citep[e.g.,][]{WO1996, Giacalone2004}.
250: %
251: However, modeling the \NL\ generation of \rel\ particles and strong
252: magnetic turbulence in collisionless shocks is
253: computationally challenging and PIC simulations will not be able
254: to fully address this problem in \nonrel\ shocks for some years to come
255: even though they can provide critical information on the plasma
256: processes producing injection that can be obtained in no other way. In
257: Appendix~\ref{estimatesforpic} we outline the requirements that a
258: PIC simulation must fulfill in order to tackle the problem of efficient
259: DSA with non-linear MFA in SNR shocks.
260:
261: In the \mc\ approximation we use here, the plasma interactions are
262: parameterized allowing us to study coupled \NL\ effects between the
263: extended shock precursor and the gas subshock. In particular, we
264: investigate the \NL\ effects caused by upstream plasma heating due to
265: magnetic field dissipation.
266:
267: The importance of the dissipation of turbulence in the shock precursor
268: can be illustrated by the following estimate. Suppose that in a shock
269: wave of speed $u_0$, the turbulence is generated by the
270: resonant cosmic ray (CR) streaming instability, so the energy density of
271: the turbulence, $U_w$, evolves approximately as
272: $u_0\,dU_w(x)/dx=v_A\,d\Pcr(x)/dx$ \citep[e.g.,][]{BL2001}, where
273: $\Pcr(x)$ is the CR pressure at position $x$ and $v_A$ is the
274: \Alf\ speed.
275: %
276: Ignoring all non-linear effects, the turbulence energy density at
277: the shock positioned at $x=0$ is $U_w(0)=\rho_0 u_0 \vazero \cdot
278: \Pcr(0)/(\rho_0 u_0^2)$.
279: %
280: The ratio $\AccEff=\Pcr(0)/(\rho_0 u_0^2)$ characterizes the
281: efficiency of acceleration and is typically assumed to be on the
282: order of ten percent or more.
283: %
284: In the above, $\rho$ is the fluid density and the subscript ``0''
285: always indicates far upstream values.
286: %
287: Suppose a fraction, $\heatpar$, of this energy goes into heating of the
288: thermal gas in the shock precursor so the energy density of the thermal
289: plasma increases by $\Delta U_{H}(0) = \heatpar U_w(0)$ at $x=0$.
290: %
291: Comparing $\Delta U_{H}(0)$ with the internal energy density of the far
292: upstream plasma, $\epsilon_0$, we find
293: %
294: \begin{equation}
295: \label{upsilonequation} \eta_H\ = \frac{\Delta U_H(0)}{\epsilon_0}
296: \approx
297: \heatpar \AccEff \frac{M_s^2}{M_A}
298: \ ,
299: \end{equation}
300: %
301: where $M_s$ is the sonic, and $M_A$ the \Alf, Mach number. If
302: $\eta_H$ is large, the thermal gas in the shock precursor is
303: strongly heated and this influences the subshock strength and the
304: particle injection efficiency. In a non-linearly modified shock, a
305: change in the injection efficiency causes the whole shock structure
306: to change.
307: For typical supernova remnant (SNR) parameters (e.g., $u_0 \sim
308: 3000$ \kmps, $T_0 \sim 10^4$ K, $n_0 \sim 0.3$ \pcc, and $B_0 \sim
309: 3$ \muG),
310: the ratio $M_s^2/M_A \approx 250$, and values of $\heatpar$ as low
311: as a few to ten percent may be important.
312:
313: Because existing analytical descriptions of MHD wave damping
314: rely on the quasi-linear approximation $\Delta B \ll B_0$, which is
315: inapplicable for strong turbulence,
316: and because an exact description of this process
317: in the framework of non-linear DSA is currently impossible
318: (see Appendix~\ref{estimatesforpic}), we propose a
319: parameterization of the turbulence damping rate.
320: In doing this, we are pursuing two goals.
321: First, we make some predictions connecting cosmic
322: ray spectra, turbulent magnetic fields
323: and plasma temperatures, which, in principle, can be tested
324: against high resolution X-ray observations in order to estimate the
325: heating of the thermal gas by turbulence dissipation.
326: And second, once heating is included in our simulation in a
327: parameterized fashion, we will be ready to implement more realistic
328: models of turbulence generation and dissipation as they are developed.
329:
330: Our \mc\ simulation can be briefly summarized as follows \citep[see][ and
331: references therein for more complete details]{EJR90,JE91,VEB2006}.
332: We describe particle transport in a plane shock by Bohm diffusion in a
333: plasma flowing in the $x$-direction with speed $u(x)$.
334: %
335: Particles move in small time steps as their local plasma frame momenta
336: are `scattered' at each step in a random walk process on a sphere in
337: momentum space.
338: %
339: Some shock heated thermal particles are injected into the
340: acceleration process when their history of random scatterings in the
341: downstream region takes them back upstream. These particles gain
342: energy and some continue to be accelerated in the first-order Fermi
343: mechanism. This form of injection is generally called `thermal
344: leakage' and was first used in the context of DSA in \citet{EJE1981}
345: \citep[see also][]{Ellison82}.
346:
347: The magnetic field determining the random walk properties
348: through the diffusion coefficient
349: is the `seed' (interstellar) magnetic field after it has been amplified
350: by a large factor by the CR streaming instability and compressed and
351: advected downstream with the flow. The streaming instability in the
352: non-linear regime ($\Delta B \gg B_0$) is described by the traditional
353: quasi-linear equations, where the instability driving term is the CR
354: pressure gradient.
355: %
356: These quasi-linear equations are extrapolated into the non-linear regime
357: in a parameterized fashion for lack of a more complete analytic
358: description. The magnetic turbulence generated by the instability is
359: assumed to dissipate at a rate proportional to the turbulence generation
360: rate, and the dissipated energy is pumped directly into the thermal
361: particle pool. An iterative scheme is employed to ensure the
362: conservation of mass, momentum, and energy fluxes, thus producing a
363: self-consistent solution of a steady-state, plane shock, with particle
364: injection and acceleration coupled to the bulk plasma flow
365: modification and to the magnetic field amplification and damping.
366:
367:
368: Our results show that even a small rate of turbulence dissipation
369: can significantly increase the precursor temperature and that this, in
370: turn, can increase the rate of injection of thermal particles. The
371: \NL\ feedback of these changes on the shock structure, however, tend
372: to cancel so that the spectrum of high energy particles is only
373: modestly affected.
374:
375: \section{Model}
376: %
377: The \mc\ simulation used here contains all of the elements of the code
378: used in \citet{VEB2006}, and previously by Ellison and co-workers, i.e.,
379: it iteratively determines a \SC, steady-state shock solution with
380: resonant \MFA. The present code, however, has been completely re-written
381: and optimized for the problem of DSA with MFA in nonrelativistic, plane
382: shocks. Besides including dissipation, the new code has been written for
383: parallel processing and can model acceleration, in a reasonable time, to
384: PeV energies\footnote{The ability of the Monte Carlo code to
385: accelerate particles to PeV energies was shown in \citet{EV2008}. The
386: results we are interested in here do not require such high energies and
387: we use a smaller dynamic range (see the note on dynamic range at the end
388: of Section~\ref{nlresults}).}.
389: %
390: The
391: particle transport is briefly described in Section
392: \ref{particletransport}, the amplification and dissipation of turbulence
393: are described in Section~\ref{amplificationanddamping}, the effects of
394: turbulence damping on the thermal plasma are described in
395: Section~\ref{plasmaheating}, and the iterative procedure used to reach a
396: self-consistent solution is given in Section~\ref{iterativeprocedure}.
397:
398: \subsection{Particle Transport and Injection in the Monte Carlo Code}
399: \label{particletransport}
400: %
401: Given a flow speed profile $u(x)$ and a diffusion coefficient $D(x,p)$
402: (which are obtained as explained below), the Monte Carlo code generates
403: a thermal distribution of particles far upstream (or close to the shock,
404: as described in Section \ref{plasmaheating} and Appendix~\ref{fastpush}),
405: and propagates them by small time steps,
406: $\delta t$, scattering their momenta with the `pitch-angle scattering'
407: scheme described in detail in \citet{EJR90}.
408: %
409: Since we assume strong turbulence ($\Delta B \gg B_0$), the concept of
410: `pitch-angle' as an angle measured with respect to the average magnetic
411: field direction loses it meaning. Instead, we define the `pitch-angle'
412: as the angle that a particle's plasma frame momentum makes with the
413: direction of the bulk flow. Each scattering is elastic and
414: isotropic in the local plasma frame, which moves at speed $u(x)$ with
415: respect to the viscous subshock located at $x=0$, and the angle of
416: scattering is random, but its maximal value is determined by $\delta t$
417: and by $D(x,p)$ \citep[see][]{EJR90}. When particles cross the subshock,
418: or move in any compressive flow, the elastic scattering in the plasma
419: frame makes them gain energy in the shock frame (see also
420: Section~\ref{mcadiabatic}).
421:
422: Injection of particles into the acceleration process occurs in the Monte
423: Carlo simulation when a formerly thermal particle first crosses the
424: viscous subshock backwards, i.e., going against the flow.
425: %
426: The number of particles that do this, and the energy they gain, are
427: determined only by the random particle histories; no parameterization of
428: the injection process is made other than the assumption of the diffusion
429: coefficient value at various particle energies.\footnote{As in previous
430: implementations of our \mc\ model, the subshock is assumed to be
431: transparent to all particles, including thermal ones, and possibly
432: important plasma effects such as a cross-shock potential or large
433: amplitude magnetic structures in the subshock layer are ignored.}
434:
435: Since microscopic plasma modeling of particle injection processes in
436: non-relativistic subshocks with an appropriate 3-D PIC code is not
437: yet available, all the models dealing with particle acceleration by
438: shocks must assume some prescription of the injection process. We
439: favor the present model for two reasons. First, our injection rate
440: is set by the scattering prescription and doesn't require any
441: additional assumptions. Second and even more important, our
442: model was shown to agree well with spacecraft observations of the
443: Earth's bow shock \citep[][]{EM87, EMP90}, interplanetary shocks
444: \citep[][]{BOEF97}, and with 1-D hybrid PIC simulations
445: \citep[][]{EGBS93}. We are aware of alternative models of particle
446: injection, such as that used by \citet{BGV2005}, in which the shock is assumed
447: transparent only to particles exceeding the thermal gyroradius by a
448: certain factor. It is possible to simulate the injection with the
449: \citet{BGV2005} recipe in the Monte Carlo code, which was done by
450: \citet{EBG2005}, but it should be noted that an injection threshold
451: may be inconsistent with the bow shock observations reported in
452: \citet{EMP90}. Our thermal injection model is also simultaneously
453: consistent with the bow shock helium and CNO observations with no
454: additional parameters or assumptions. A comparative analysis of the
455: different injection recipes is beyond the scope of our present study.
456:
457: A peculiarity of our approach is that in order to separate the CR
458: particles from the thermal ones we use their history, and not their
459: energy. By our definition, a thermal particle is one that we had
460: introduced into the simulation upstream with a random thermal energy and
461: that may have crossed the subshock going downstream, but has never crossed
462: it back. Once a particle crosses the subshock (the coordinate $x=0$, to
463: be more precise) in the upstream direction, it by our definition
464: is injected and becomes a CR particle
465: (see also Appendix~\ref{special_th_vs_cr}).
466:
467: As particles are propagated and scattered, their contributions to fluxes
468: of mass, momentum, and energy are calculated at select positions. We
469: also calculate the pressure produced by the thermal particles $\Pth(x)$
470: and the spectrum of pressure of the CR particles $\Pcrhat(x,p)$.
471: (see details in Appendix~\ref{fastpush}). The latter is
472: related to the total CR pressure $\Pcr(x)$ as
473: %
474: \begin{equation}
475: \Pcr(x) = \int\limits_0^{\infty}{\Pcrhat(x,p)dp}
476: \ ,\end{equation}
477: %
478: and $\Pcrhat(x,p)$ is then used to calculate the magnetic field
479: amplification and dissipation, as described in Section
480: \ref{amplificationanddamping}.\footnote{We note that notation has
481: changed slightly from that used in \citet{VEB2006}. The subscript `tot'
482: indicating integral quantities has been eliminated, and these quantities
483: are now denoted as $\Pcr(x)$, $L(x)$, etc. The quantities representing
484: spectral densities of pressure, flux, etc., are denoted with the same
485: letters, but a different set of arguments: $\Pcr(x,p)$, $L(x,k)$. To
486: avoid ambiguity, explicit definitions relating the spectral densities to
487: the integral densities are provided throughout the text. We
488: also dropped the bar above the dissipation rate term $L(x)$
489: for simplicity.}
490:
491: We assume in this paper that the acceleration is size-limited, and model
492: the finite size of the shock with a free escape boundary (FEB), located
493: at position $\xFEB<0$ far upstream of the shock. All CR particles
494: crossing the boundary escape freely from the system. The maximal
495: particle momentum, $\pmax$, is thus determined when
496: the upstream diffusion length becomes comparable to $|\xFEB|$. For a
497: spherical SNR blast wave, $|\xFEB|$ is on the order of the radius of the
498: remnant.
499:
500:
501: \subsection{Turbulence Amplification and Dissipation}
502:
503: \label{amplificationanddamping}
504:
505: We assume that far upstream there exists a uniform magnetic field,
506: $B_0$, parallel to the flow direction which is perturbed by transverse
507: \Alf ic fluctuations with a power law energy spectrum. It is further
508: assumed that these fluctuations produce Bohm diffusion for particles
509: of all energies.
510: %
511: Closer to the shock, where a population of accelerated particles
512: drifting upstream is present, this seed turbulence is amplified by the
513: CR streaming instability and additionally strengthened by the plasma
514: compression. Once amplification begins, the spectrum of turbulence is no
515: longer restricted to any particular form\footnote{We assumed
516: $U_{\pm}\propto k^{-1}$ for the seed turbulence. The choice of the seed
517: turbulence spectrum does not significantly affect our results
518: for two reasons. First, the diffusion coefficient assumed here only
519: depends on the total power in the turbulence and is insensitive to the
520: shape of the spectrum, and second, the rapidly growing fluctuations due
521: to the streaming instability quickly overpower the seed
522: spectrum. Despite our using the Bohm diffusion coefficient, we still
523: keep track of the turbulence spectrum in the simulation since this
524: will be used in future work where the diffusion coefficient is
525: determined self-consistently from the wave spectrum.}.
526: Assuming that the turbulence is described by the
527: quantities $U_-(x,k)$ and $U_+(x,k)$ ($k$ is the wavenumber of
528: turbulence harmonics, $U_-$ and $U_+$ are the spectra of energy
529: density of structures propagating in the upstream and the downstream
530: directions with respect to the thermal plasma, respectively), we model
531: the evolution of the turbulence, as it is being advected with the plasma
532: and amplified, with the following equations:
533: %
534: \begin{equation}
535: \label{ampeq}
536: E_{\pm}[U] = (1-\heatpar)G_{\pm}[U] + I_{\pm}[U].
537: \end{equation}
538: %
539: Here, for readability, we abbreviated as $E$ the evolution operator,
540: as $G$ the growth operator and as $I$ the wave-wave interactions
541: operator, acting on the spectrum of turbulence energy density
542: $U=\{U_-(x,k), U_+(x,k)\}$. These quantities are defined as follows:
543: %
544: \begin{eqnarray}
545: \nonumber
546: E_{\pm}[U] & = &
547: \left(u \pm V_G\right)\frac{\partial}{\partial x} U_{\pm} + \qquad \\
548: && \qquad U_{\pm}\frac{d}{dx}\left(\frac32 u \pm V_G\right), \\
549: \nonumber
550: G_{\pm}[U] & = &
551: \mp\frac{U_\pm}{U_+ + U_-} V_G \times \qquad \\
552: && \qquad \frac{\partial \Pcrhat(x,p)}{\partial x}
553: \left| \frac{dp}{dk} \right|, \\
554: I_{\pm}[U] & = &
555: {\pm} \frac{V_G}{\rgzero}\left(U_- - U_+\right).
556: \end{eqnarray}
557: %
558: Here and throughout the paper, $\rgzero=m_p u_0 c / (eB_0)$ -- the
559: gyroradius of a particle with a plasma frama speed equal to $u_0$
560: in the far upstream magnetic field $B_0$.
561: The parameter $\heatpar$ describes the turbulence dissipation rate, and
562: for $\heatpar=0$, equation~(\ref{ampeq}) is exactly
563: what was used in \citet{VEB2006}, except there it was
564: assumed that $V_G\ll u(x)$. Keeping $V_G$ relative to $u(x)$ results in
565: a slightly greater amplification of magnetic field.
566: %
567: In this system $u=u(x)$ is the flow speed and
568: $V_G=V_G(x)$ is the parameter defining the turbulence growth rate
569: and the wave speed\footnote{As explained in \citet{VEB2006},
570: in the quasilinear case, $\Delta B \ll B_0$,
571: the wave speed and the speed determining turbulence growth rate are
572: both equal to the \Alf\ speed,
573: $V_G (x) = v_A = B_0 / \sqrt{4 \pi \rho(x)}$.
574: In the case of strong turbulence, $\Delta B \gtrsim B_0$,
575: we hypothesise that the resonant streaming instability can
576: still be described by equations~(\ref{ampeq}) with $V_G$
577: being a free parameter ranging from $B_0 / \sqrt{4 \pi \rho(x)}$
578: to $\Beff / \sqrt{4 \pi \rho(x)}$.}.
579: %
580:
581: For simplicity, we assume $V_G (x) = B_0 /\sqrt{4 \pi \rho(x)}$,
582: in the present work, corresponding to $\fAlf=0$ in \citet{VEB2006},
583: and emphasize that the use of equation~(\ref{ampeq}) to
584: describe the streaming instability when $\Delta B \gtrsim B_0$ is only a
585: parameterization.
586: %
587:
588: The growth operator $G$, which describes the turbulence amplification by
589: the CR streaming instability, is proportional to the gradient of
590: $\Pcr(x,p)$, the latter being the spectrum of pressure of the CR
591: particles driving the instability. $\Pcr(x,p)$ is
592: computed in the Monte Carlo simulation from the trajectories of
593: particles, and the momentum $p$ at which it is taken in (\ref{ampeq}) is
594: the momentum resonant with the turbulent structures with wavenumber
595: $k$. The resonance condition assumed is
596: %
597: \begin{equation}
598: k \frac{c p}{e B_0} = 1.
599: \end{equation}
600: %
601:
602: The parameter $\heatpar$ enters the
603: equations of turbulence evolution (\ref{ampeq}) through the factor
604: $(1-\heatpar)$, by which the term $G$ describing the magnetic field
605: amplification by the CR streaming is reduced (it is assumed that $0 \leq
606: \heatpar \leq 1$). By writing this, we assumed simply that at all
607: wavelengths only a fraction $(1 -\heatpar)$ of the
608: instability growth rate goes into the magnetic turbulence, and the
609: remaining fraction $\heatpar$ is lost in the dissipation process.
610: %
611: The factor $(1-\heatpar)$ in (\ref{ampeq}) can also be understood in the
612: following way: \citet{VEB2006} derived their equations~(11) and (12) from
613: equation (3) by assuming that the loss term $L=0$ ($L$ here is $\bar{L}$ in
614: their notation), but now we
615: are assuming that
616: \begin{equation}
617: \label{dampingparameterization_spc}
618: L=- \heatpar \cdot v_\mathrm{a, x}\frac{\partial \Pcr}{\partial x} .
619: \end{equation}
620: For $\heatpar=0$ no dissipation occurs, and the
621: CR streaming instability pumps the energy of the accelerated particles
622: into the magnetic turbulence, amplifying the effective magnetic field
623: most efficiently. For $\heatpar=1$
624: the additional turbulence energy produced by the instability is assumed
625: to immediately dissipate, and the scattering in the shock is assumed to
626: be provided only by the seed turbulence slightly increased by the plasma
627: compression.
628:
629: The energy dissipation rate at all wavelengths is then
630: %
631: \begin{equation}
632: \label{dampingparameterization}
633: L(x)=\int\limits_0^{\infty} L(x,k)\: dk = \heatpar \cdot V_G \frac{d
634: \Pcr(x)}{dx},
635: \end{equation}
636: %
637: and we assume that all this energy goes directly into heating
638: thermal particles. The
639: modeling of the thermal plasma heating is covered in detail in Section
640: \ref{plasmaheating}.
641:
642: When equation~(\ref{ampeq}) is solved, the resulting
643: $U_{\pm}(x,k)$ are used to calculate the amplified effective magnetic
644: field
645: \begin{eqnarray}
646: \nonumber
647: \Beff(x) &=& \left[ 4 \pi \int\limits_0^{\infty} U_-(x,k)\:dk +\qquad \right. \\
648: \label{beffdef}
649: && \left. 4\pi \int\limits_0^{\infty} U_+(x,k) \: dk \right]^{1/2},
650: \end{eqnarray}
651: the turbulence pressure
652: \begin{equation}
653: \label{pwdef}
654: P_w(x) = \frac{\Beff^2(x)}{8 \pi}
655: \end{equation}
656: and the turbulence energy flux
657: \begin{eqnarray}
658: \nonumber
659: F_w(x) &=& \left( \frac32 u - V_G \right) \int\limits_0^{\infty}U_-(x,k)\:dk+\qquad\\
660: \label{fwdef}
661: &&\left( \frac32 u + V_G \right) \int\limits_0^{\infty}U_+(x,k)\:dk
662: \end{eqnarray}
663: (see equations (15), (18) and (19) in \citet{VEB2006}), which are then
664: used in the derivation of a self-consistent solution, as discussed in
665: Section~\ref{iterativeprocedure} below. In the present paper we do
666: not neglect $V_G$ in the sum with $u$, which results in a slightly lower
667: compression ratio $\rtot$ than in the case $3/2 u \pm V_G \approx 3/2
668: u$, that was adopted in \citet{VEB2006}. Note that the factor
669: $3/2$ is valid for \Alf\ wave-like modes, which is implicitly assumed
670: by our using the system of equations (\ref{ampeq}).
671:
672: As mentioned by \citet{CBAV2008}, effects from the transmission and
673: reflection of \Alf\ waves at the subshock could be important, and
674: accounting for these effects may further lower the compression ratio
675: $\rtot$.
676: %
677: We do not account for reflection and linear transformations of \Alf\
678: waves to other MHD modes (magnetosonic, entropy, etc.) at the subshock
679: (see McKenzie and Westphal 1970) in the present paper, because a correct
680: description of these effects on the subshock in highly turbulent media
681: must contain simultaneously some other comparable effects.
682: %
683: Indeed, MHD waves interacting with a shock produce stochastic ripples in
684: the shock surface and these ripples produce an effective broadening of
685: the shock spatial structure determined by the turbulence spectrum
686: \citep[see][]{Bykov1982}.
687: %
688: Moreover, as we argue in Appendix~\ref{special_th_vs_cr}, accounting for
689: suprathermal particles modifies the standard Rankine-Hugoniot relations
690: at the subshock and these effects could make it impossible to identify
691: the subshock as a plane discontinuity, as assumed in all existing
692: semi-analytic models.
693: %
694: Clearly, these phenomena are important and require further investigation
695: but they are beyond the scope of our simplified description of
696: MFA. Nevertheless, we believe our predictions regarding the turbulence
697: dissipation in the shock precursor are qualitatively correct.
698:
699:
700:
701: \subsection{Heating of Thermal Plasma}
702: \label{plasmaheating}
703: %
704: Repeating the derivation of equation (9) in \citet{MV82}, one
705: obtains for a steady-state shock:
706: %
707: \begin{equation}
708: \label{pressuregrowth}
709: \frac{ u\rho^{\gamma}}{\gamma-1}
710: \frac{d}{dx}\left(\Pth \rho^{-\gamma} \right) = \Lbar(x)
711: \ .
712: \end{equation}
713: %
714: Here and elsewhere, $\rho=\rho(x)$, $u=u(x)$, and $\Pth=\Pth(x)$.
715: The quantity $\Lbar(x)$ is the dissipation rate defined in
716: (\ref{dampingparameterization}), and the ratio of specific
717: heats of an ideal nonrelativistic gas is $\gamma=5/3$.
718: %
719: For $\Lbar(x)=0$, equation (\ref{pressuregrowth})
720: reduces to the adiabatic heating law,
721: $\Pth \sim \rho^{\gamma}$ and, for a non-zero $\Lbar(x)$, it describes
722: the heating of the thermal plasma in the shock precursor due to the
723: dissipation of magnetic turbulence.
724: %
725: The fluid description of heating given by
726: equation~(\ref{pressuregrowth}), while it doesn't include details of
727: individual particle scattering, can be used in the \mc\ simulation to
728: replace particle scattering and
729: determine heating in the shock precursor. This merging of analytic and
730: \mc\ techniques, or \FP\ (\FPshort), is described in detail
731: in Appendix \ref{fastpush}.
732:
733: The \FPshort\ only affects our treatment of thermal
734: (i.e., not injected)
735: particles in the precursor and involves two steps. The first is to
736: introduce thermal
737: particles into the simulation, not far upstream as we did before, but at
738: some position $\xFP < 0$ close to the subshock and with a temperature
739: equal to what Eq.~(\ref{pressuregrowth}) and the ideal gas law
740: suggest:
741: %
742: \begin{equation}
743: \label{idealgaslaw}
744: T(x)=\frac{\Pth(x)}{k_B n_0 (u_0/u(x))},
745: \end{equation}
746: %
747: where all quantities are taken at $x=\xFP$ and $\Pth$ is calculated from
748: (\ref{pressuregrowth}) ($n_0$ and $u_0$ are the far upstream number
749: density and shock speed, and $k_B$ is
750: Boltzmann's constant). The second step is to calculate the
751: thermal gas pressure throughout the precursor (at $x<\xFP$) using
752: (\ref{pressuregrowth}) instead of tracing particle motions.
753:
754: The main effects of turbulence dissipation in our model are:
755: %
756: \newlistroman\listromanDE a decrease in the value of the amplified field
757: $\Beff(x)$, which determines the diffusion coefficient, $D(x,p)$;
758: %
759: \listromanDE an increase in the temperature of particles just upstream of
760: the subshock, which influences the injection of particles into the
761: acceleration process, and
762: %
763: \listromanDE
764: an increase in the thermal particle pressure $\Pth(x<0)$,
765: and a decrease in the turbulence pressure $P_w(x)$,
766: which enter the conservation equations
767: described in Section \ref{iterativeprocedure}.
768: Since all of these processes are coupled, a change in dissipation
769: influences the overall structure of the shock.
770:
771: \subsection{Closure of the Model}
772: \label{iterativeprocedure}
773: %
774: Typically, we begin our simulation by propagating particles, with a
775: pre-set diffusion coefficient $D(x,p)$, in an unmodified shock, where the
776: flow speed jumps discontinuously from $u_0$ to $u_2$, that is,
777: $u(x)=u_0$ for $x<0$ and $u(x)=u_2$ for $x>0$.\footnote{Everywhere in
778: the text, unless otherwise noted,
779: the subscript `0' indicates a far upstream value, `1'
780: indicates a value just upstream of the subshock, and `2' indicates a
781: downstream value. For example, $u_0=u(x=-\infty)$, $u_1=u(\xtr)$, and
782: $u_2=u(x>0)$. See Appendix~\ref{mcadiabatic} for definition of $\xtr$.}
783: This allows us to calculate the various fluxes and other quantities,
784: such as the CR pressure spectrum $\Pcrhat(x,p)$, at any position $x$.
785: The latter is used to solve Eq.~(\ref{ampeq}), which yields the
786: turbulence spectra $U_{\pm}(x,k)$ and, subsequently, the amplified
787: effective field $\Beff(x)$ and the pressure of the magnetic
788: turbulence $P_w(x)$. The spectrum $\Pcrhat(x,p)$ also provides
789: the turbulence dissipation rate (\ref{dampingparameterization})
790: and the resulting pressure of the turbulence-heated gas
791: (\ref{pressuregrowth}).
792:
793: The equations for the conservation of mass and momentum fluxes are:
794: %
795: \begin{equation}
796: \label{fluxmass}
797: \rho(x) u(x) = \rho_0 u_0
798: \end{equation}
799: %
800: ($\rho$ and $u$ are the mass density and the flow speed) and
801: %
802: \begin{equation}
803: \label{fluxmomentum}
804: \momentumflux(x) = \upstreammomentumflux,
805: \end{equation}
806: %
807: where
808: $\momentumflux(x)$ is the flux of the $x$-component of momentum in the
809: $x$-direction including the contributions from particles and turbulence,
810: and $\upstreammomentumflux$ is the far upstream value of momentum flux, i.e.,
811: \begin{equation}
812: \upstreammomentumflux =
813: \rho_0 u_0^2 + \Pthzero + \Pwzero
814: \ .
815: \end{equation}
816: The quantity $\momentumflux$ is defined as
817: \begin{equation}
818: \label{mfasmoment}
819: \momentumflux(x) = \int p_x v_x f(x, \pvector) d^3p + P_w(x),
820: \end{equation}
821: (here $p_x$ and $v_x$ are the $x$-components of momentum and
822: velocity of particles, and $f(\pvector)$ is their distribution
823: function, all measured in the shock frame),
824: and in the simulation it is calculated by summing the contributions
825: of the particles crossing a grid location and adding the turbulence
826: pressure $P_w$ defined in (\ref{pwdef}). See details of this computation in
827: Appendix~\ref{fastpush}.
828:
829: Initially, in our simulation, the shock doesn't have a \SC\
830: structure because we start with an unmodified shock and
831: $\momentumflux(x)$ is overestimated at all locations where
832: accelerated particles are present.
833: The next step is to use the calculated macroscopic quantities to find a
834: new $u(x)$ that reduces the mismatch between the local momentum
835: flux and the far upstream value of it $\upstreammomentumflux$
836: for $x<0$. We do this by calculating
837: \begin{eqnarray}
838: \label{iteration}
839: u'(x)=u(x) + s \cdot \frac{\momentumflux(x) - \upstreammomentumflux}{\rho_0 u_0},
840: \end{eqnarray}
841: where $u'(x)$ is the predicted flow speed for the
842: next iteration, and $s$ is a small positive number (typically around
843: 0.1), characterizing the pace of the iterative procedure. At this
844: point we also refine our
845: estimate for the particle diffusion coefficient which, as in
846: \citet{VEB2006}, is assumed to be Bohm diffusion such that the particle
847: mean free path is equal to its gyroradius in the effective, amplified
848: field $\Beff(x)$:
849: %
850: \begin{equation}
851: \label{bohm}
852: D(x,p)=\frac{v\lambda}{3} = \frac{vcp}{3e\Beff(x)},
853: \end{equation}
854: %
855: where $\Beff(x)$ is defined in (\ref{beffdef}).
856: %
857:
858: The predicted $u(x)$ and $D(x,p)$ are then used in a new iteration where
859: particles are injected and propagated. The calculated CR pressure, momentum
860: flux, etc. are then used to refine the guesses for $u(x)$ and $D(x,p)$
861: for the next iteration, and so on. This procedure is continued until all
862: quantities converge.
863:
864: In order to conserve momentum and energy, the compression ratio,
865: $\rtot=u_0/u_2$, must be determined self-consistently with the
866: shock structure. To determine $\rtot$ we use the condition of the
867: conservation of energy flux given by:
868: %
869: \begin{equation}
870: \label{fluxenergy}
871: \energyflux(x) + \Qesc(x) = \upstreamenergyflux,
872: \end{equation}
873: %
874: where $\energyflux(x)$ is the energy flux of particles and turbulence in
875: the $x$-direction, $\Qesc$ is the energy flux of escaping particles
876: at the FEB,\footnote{Particle escape at an upstream FEB also
877: causes the mass and momentum fluxes to change but these changes are
878: negligible as long as $u_0 \ll c$ \citep[][]{Ellison85}.} and the far
879: upstream value of the energy flux is
880: \begin{equation}
881: \upstreamenergyflux = \frac{1}{2} \rho_0 u_0^3 +
882: \frac{\gamma}{\gamma-1}\Pthzero u_0 + \Fwzero.
883: \end{equation}
884: The quantity $\energyflux(x)$ is defined as
885: \begin{equation}
886: \label{efasmoment}
887: \energyflux(x) = \int K v_x f(x,\pvector) d^3 p + F_w(x),
888: \end{equation}
889: $K$ being the kinetic energy of a particle with momentum $p$ measured in
890: the shock frame, and $F_w$ is the energy flux of the turbulence
891: defined in (\ref{fwdef}). The details of calculating
892: $\energyflux(x)$ in the simulation are given
893: in Appendix~\ref{fastpush}, and the explanation of
894: how $\energyflux(x)$ is used in an iterative procedure converging to a
895: consistent $\rtot$ is given in Appendix \ref{compressionratio}.
896:
897:
898:
899:
900:
901: \section{Results}
902: %
903: \subsection{Particle Injection in Unmodified Shocks (Subshock Modeling)}
904: \label{injectionvsdissipation}
905:
906: In order to isolate the effects of plasma heating on particle injection,
907: we first show results for unmodified shocks, i.e., $u(x<0)=u_0$ and
908: $u(x>0)=u_0/\rtot$, with fixed $\rtot$. In these models particle
909: acceleration, \MFA\ and turbulence damping are included consistently
910: with each other, but we do not obtain fully \SC\ solutions
911: conserving momentum and energy, since this requires the shock
912: to be smoothed, while we intentionally fix $u(x)$.
913:
914: %\clearpage
915: \begin{figure}
916: \epsscale{1.0}
917: \plotone{f1.eps}
918: \caption{Dependence of magnetic field amplification and particle
919: injection on the rate of magnetic turbulence dissipation in unmodified
920: shocks. Results for shocks with compression ratios $\rtot$=3.0, 3.2,
921: 3.4, 3.5 and 3.6 are represented by the thin solid, dotted, dashed,
922: dash-dotted and thick solid lines, respectively. The $x$-axis variable
923: is the turbulence dissipation rate $\heatpar$ (constant throughout a
924: shock), and the plotted quantities are the amplified downstream
925: effective magnetic field $\Befftwo$, the subshock sonic Mach number
926: $\Msone$, the fraction of simulation particles
927: injected into the acceleration process $\fcr$, and the ratio of
928: amplified field to $\Btrend$ downstream from the shock.
929: \label{fig_unmod_inj}}
930: \end{figure}
931: %\clearpage
932:
933: In Fig.~\ref{fig_unmod_inj} we show results where the compression ratio is
934: varied between $\Rtot=3$ and $3.6$ as indicated. In all models, $u_0 =
935: 3000$ \kmps, $T_0 = 10^4$ K, $n_0 = 0.3$\,\pcc\ and $B_0 = 3$\,\muG\
936: (the corresponding sonic and \Alf\ Mach numbers are $\Mszero \approx
937: \Mazero \approx 250$). The FEB was set at $\xFEB = -3\cdot 10^4\:
938: \rgzero$ (our spatial scale unit $\rgzero=m_p u_0 c / (eB_0)$),
939: and for each $\Rtot$ we obtained results for different values
940: of $\heatpar$ between 0 and 1. The values plotted in the
941: top three panels of Fig.~\ref{fig_unmod_inj}
942: are the amplified magnetic field downstream, $\Befftwo$, the Mach number
943: right before the shock, $\Msone$ (this is not equal to $\Mszero$ because
944: of the plasma heating due to turbulence dissipation), and the fraction
945: of thermal particles in the simulation that crossed the shock in the
946: upstream direction at least once (i.e., got injected), $\fcr$. The bottom
947: panel shows the ratio of the calculated downstream effective magnetic
948: field $\Befftwo$ to trend values $\Btrend(\heatpar)$;
949: what is meant by ``trend'' is explained below.
950:
951: Looking at the curve for $\Befftwo$ in the $\rtot=3.0$ model, one sees
952: an easy to explain behavior: as the magnetic turbulence dissipation rate
953: increases, the value of the amplified magnetic field decreases, going
954: down to $B_0 \rtot^{3/4}$ (the upstream field compressed at the shock)
955: for $\heatpar=1$. Increasing $\heatpar$ simply causes more energy to be
956: removed from magnetic turbulence and put into thermal particles, thus
957: decreasing the value of $\Befftwo$.\footnote{We use this result, with
958: well understood behavior, to test the implementation of turbulence
959: dissipation in our simulation.}
960: %
961: In our model, the amount of dissipated turbulence energy scales linearly
962: with $\heatpar$. Therefore, the trend of $\Befftwo$ changing with
963: $\heatpar$ under the assumption that the total efficiency of the
964: streaming instability is unchanged, but the energy is channelled
965: from magnetic turbulence to thermal particles, can be described as
966: \begin{eqnarray}
967: \label{btrend}
968: \nonumber
969: \Btrend^2(\heatpar) &=&
970: \left(B_0 \rtot^{3/4}\right)^2 + (1-\heatpar) \times \qquad\quad\\
971: \label{btwotrend}
972: &&
973: \left[ \left. \Befftwo^2 \Big.\right|_{\heatpar=0}
974: - \left(B_0 \rtot^{3/4}\right)^2 \right],
975: \end{eqnarray}
976: where the first term on the right hand side is the $B_0$
977: compressed at the shock \citep[the compression factor $\rtot^{3/4}$
978: is explained by equation (13) in][]{VEB2006}, and the
979: second term is proportional to the amount of the magnetic
980: turbulence energy density generated by the instability for $\heatpar=0$,
981: reduced by the factor $(1-\heatpar)$. Neglecting $B_0$,
982: Eq.~(\ref{btwotrend}) predicts $\Btrend \propto
983: \sqrt{1-\heatpar}$. The comparison of $\Befftwo$ with
984: $\Btrend(\heatpar)$ from (\ref{btwotrend}) is shown in the bottom panel
985: of Fig.~\ref{fig_unmod_inj}. It is clear that the
986: above trend matches the calculations very well for $\rtot=3.0$.
987: %
988: One also can see that the sonic Mach
989: number in this simulation decreased from the upstream value of $250$ to
990: approximately $20$, and that the fraction of injected particles remained
991: almost constant as $\heatpar$ was varied.
992:
993: The curves for $\rtot=3.2$ demonstrate the same behavior, and the shape
994: of the $\Befftwo$ curve is similar to the one for $\rtot=3.0$, with the
995: higher compression ratio producing a higher value of the amplified
996: magnetic field due to a greater number of particles getting
997: injected. The calculated $\Befftwo$ deviates from the trend
998: $\Btrend(\heatpar)$ more than in the $\rtot=3.0$ case, and this
999: deviation marks the emergence of an effect that becomes more
1000: pronounced as $\rtot$increases.
1001:
1002: The plots for $\rtot \gtrsim 3.4$ present a
1003: qualitatively different behavior from those with $\rtot \lesssim
1004: 3.2$. The downstream magnetic field $\Befftwo$ does
1005: decrease with increasing $\heatpar$, but not as rapidly as in the
1006: previous two cases, and there is a switching point at $\heatpar \approx
1007: 0.95$ in the curves for $\Msone$ and $\fcr$. The bottom panel of
1008: Fig.~\ref{fig_unmod_inj} shows a deviation of $\Befftwo$ from the trend
1009: (\ref{btwotrend}) by a large factor in the $\rtot = 3.4$ case. This
1010: effect becomes even more dramatic for $\rtot=3.5$ and $\rtot=3.6$ where
1011: $\Befftwo$, contrary to expectations, increases with $\heatpar$ before
1012: $\heatpar\rightarrow 1$. The fact that the final energy in turbulence
1013: can increase as more energy is transferred from the turbulence to heat
1014: indicates the \NL\ behavior of the system and shows how sensitive the
1015: acceleration is to precursor heating.
1016:
1017: Indeed, if the upstream plasma is not heated sufficiently (as in the
1018: $\rtot < 3.4$ cases), then the thermal particles first reaching the
1019: shock are cold ($T_1 \sim T_0$), and just upstream of the subshock
1020: the sonic Mach number $\Msone$ has a large value (much greater than $10$
1021: for any $\heatpar$). As these thermal particles cross the shock, their
1022: momenta in the downstream plasma frame lie in a very narrow cone opening
1023: in the downstream direction, with the opening angle around $\theta \sim
1024: \Msone^{-1}$, making it equally difficult for most particles to turn
1025: around, cross the subshock backwards and get injected into the
1026: acceleration process. As long as $\theta$ is small (or $\Msone$ is
1027: large enough), the number of injected particles is insensitive to the
1028: exact opening angle and $\fcr$ stays relatively constant, as seen in the
1029: third panel of Fig.~\ref{fig_unmod_inj} for the $\rtot=3.0$ case.
1030:
1031: The injection rate increases with $\Rtot$ and for $\Rtot=3.6$, even with
1032: $\heatpar$ as low as 0.1, the CR-generated turbulence heats the thermal
1033: plasma through dissipation enough to lower $\Msone$ to $\sim 10$.
1034: Increasing $\heatpar$ further lowers $\Msone$ even more, quickly
1035: increasing the probability that a downstream thermal particle will
1036: return upstream, thus boosting the injection rate ($\fcr$ rapidly goes
1037: up with $\heatpar$ in the $\rtot \geq 3.4$ cases). It turns out that
1038: this effect overcomes the reduction of $\Befftwo$ due to damping, and
1039: $\Befftwo$ starts increasing with $\heatpar$.
1040:
1041: For any $\rtot$, at some high enough value of $\heatpar$ (near $0.90$),
1042: the decrease in magnetic turbulence due to dissipation dominates the
1043: increase in injection and the magnetic field drops. As $\heatpar
1044: \rightarrow 1$, and $\Befftwo$ becomes small enough, the efficiency of
1045: particle acceleration reduces sufficiently to cut down the total energy
1046: put into magnetic turbulence and the corresponding fraction of this energy
1047: dissipated into the thermal particles. This causes $\Msone$ to turn up
1048: and $\fcr$ to turn down at about $\heatpar=0.95$,
1049: as shown in the Fig.~\ref{fig_unmod_inj}.
1050:
1051: It is worth mentioning that the observed increase of particle
1052: injection due to the precursor plasma heating is a consequence of the
1053: thermal leakage model of particle injection adopted here. In this model,
1054: a downstream particle, thermal or otherwise, with plasma frame speed
1055: $v>u_2$, has a probability to return upstream which increases with $v$
1056: \citep[see][for a discussion of the probability of returning
1057: particles]{Bell78a}. That is, we assume that the subshock is transparent
1058: to all particles with $v>u_2$, but only some of these particles get
1059: injected, depending on their random histories. Particles that don't get
1060: injected are convected far downstream out of the system. An alternative
1061: model of injection \citep[see, for example,][]{BGV2005} is one where
1062: only particles with a gyroradius greater than the shock thickness can
1063: get injected. The assumption with this threshold injection model is that
1064: the subshock thickness is comparable to the gyroradius of a downstream
1065: thermal particle and only those particles with speeds $v > \xi
1066: v_\mathrm{th}$ can be injected. Particles with $v < \xi v_\mathrm{th}$
1067: are somehow blocked by the subshock. Here, $v_\mathrm{th}$ is the
1068: downstream thermal speed and $\xi$ is a free parameter, typically taken
1069: to be between 2 and 4.
1070:
1071: Despite being similar, it can be expected that these injection models
1072: will react differently to precursor heating.
1073: Namely, in the \citet{BGV2005} model the fraction of
1074: injected particles may be insensitive to the precursor heating if $\xi$
1075: is fixed\footnote{\citet{AB2006}, who performed a brief analysis of the
1076: impact of the turbulence dissipation on the \NL\ shock structure in a
1077: way similar to ours, did not report an influence of heating on the
1078: injection rate.}, because the same particle injection rate occurs
1079: regardless of the pre-subshock temperature $T_1$ and downstream
1080: temperature $T_2$.
1081: %
1082: While both models are highly simplified
1083: descriptions of the complex subshock
1084: \citep[see, e.g.][]{Malkov98, GE2000}, they offer two scenarios for
1085: grasping a qualitatively correct behavior of a shock where
1086: particle injection and acceleration are coupled
1087: to turbulence generation and flow modification. Hopefully, a
1088: clearer view of particle injection by self-generated turbulence in a
1089: strongly magnetized subshock will become available when
1090: relevant full particle PIC or hybrid
1091: simulations are performed.
1092:
1093: With the general trends observed here in mind,
1094: we now show how \NL\ effects modify the effect dissipation has
1095: on injection and MFA.
1096:
1097:
1098: \subsection{Fully Nonlinear Model}
1099:
1100: \label{nlresults}
1101:
1102: In this section we demonstrate the results of the fully nonlinear
1103: models, in which the flow structure, compression ratio, magnetic
1104: turbulence, and particle distribution are all determined \SCly, so
1105: that the fluxes of mass, momentum and energy are conserved across the
1106: shock.
1107:
1108: We use two sets of parameters, one with the far upstream gas
1109: temperature $T_0=10^4$~K and the far upstream particle density $n_0 =
1110: 0.3$~\pcc, typical of the cold interstellar medium (ISM), and one with
1111: $T_0=10^6$ K and $n_0=0.003$ \pcc, typical of the hot ISM.
1112: %
1113: In both cases we assumed the shock speed $u_0 = 5000$ \kmps, and the
1114: initial magnetic field $B_0 = 3$ \muG\ (giving an equipartition of
1115: magnetic and thermal energy far upstream, $n_0 k_B T_0 \approx
1116: B_0^2/(8\pi)$). The corresponding sonic and \Alf\ Mach numbers are $M_s
1117: \approx M_A \approx 400$ in both cases). The size of the shocks
1118: was limited by a FEB located at $\xFEB = -10^5 \rgzero\approx -3 \cdot
1119: 10^{-4}$ pc.
1120: %
1121: For both cases, we ran seven
1122: simulations with different values of the dissipation rate $\heatpar$,
1123: namely $\heatpar \in \{ 0; 0.1; 0.25; 0.5; 0.75; 0.9;
1124: 1.0\}$. Also, for the \hotism\ case we ran a
1125: simulation neglecting the streaming instability effects, i.e., keeping
1126: the magnetic field constant throughout the shock and assuming that the
1127: precursor plasma is heated only by adiabatic compression (this model
1128: will be referred to as the `no MFA case').
1129:
1130: \clearpage
1131: \begin{deluxetable}{ lrrrrrrr }
1132: \tabletypesize{\small}
1133: \tablecaption{Summary of Non-linear Simulation in a Cold ISM \label{sumnl_cold}}
1134: \tablewidth{0pt}
1135: \tablehead{
1136: $\heatpar$ & 0.00 & 0.10 & 0.25 & 0.50 & 0.75 & 0.95 & 1.00
1137: }
1138: \startdata
1139: $\rtot$ & 16.0 & 16.2 & 14.5 & 14.6 & 14.0 & 13.2 & 13.0 \\
1140: $\rsub$ & 2.95 & 2.83 & 2.75 & 2.59 & 2.50 & 2.50 & 2.51 \\
1141: $\Befftwo$, \muG & 345 & 323 & 284 & 232 & 158 & 71 & 21 \\
1142: $\Btrend$, \muG & 345 & 327 & 299 & 245 & 174 & 80 & 21 \\
1143: $\left<T(x<0)\right>$, $10^4$ K
1144: & 1.06 & 4.3 & 9.0 & 17 & 26 & 37 & 56 \\
1145: $T_1$, $10^4$ K & 3.3 & 68 & 160 & 330 & 490 & 610 & 650 \\
1146: $T_2$, $10^4$ K & 1400 & 1500 & 1600 & 1600 & 1800 & 2000 & 2200 \\
1147: $\Msone$ & 44 & 9.5 & 6.3 & 4.2 & 3.5 & 3.3 & 3.2 \\
1148: $\fcr$, \% & 1.0 & 1.2 & 1.6 & 2.1 & 2.6 & 3.1 & 3.2 \\
1149: $\pmax / m_pc$ & 500 & 450 & 400 & 350 & 250 & 150 & 80 \\
1150: $\left<\gamma(x<0)\right>$
1151: & 1.33 & 1.33 & 1.33 & 1.33 & 1.34 & 1.34 & 1.34 \\
1152: $\gamma_2$ & 1.38 & 1.38 & 1.38 & 1.39 & 1.39 & 1.40 & 1.41 \\
1153: $\xtr / \rgzero$ & -0.005&-0.001 & -0.001& -0.001&-0.002 & -0.004& -0.02 \\
1154: $\xFP / \rgzero$ & -0.04 &-0.06 & -0.06 & -0.09 & -0.23 &-0.57 & -2.1 \\
1155: \enddata
1156: \footnotesize\tablecomments{Here $\heatpar$, the fraction of dissipated energy, is the
1157: model input parameter, and the rest are results of the self-consistent
1158: simulation, as follows: $\rtot=u_0/u_2$ is the total shock compression ratio,
1159: $\rsub=u_1/u_2$ is the subshock compression, $\Befftwo$ is the amplified
1160: effective magnetic field downstream, $\Btrend$ is the trend value
1161: calculated from Eq.~(\ref{btrend}), $\left<T(x<0)\right>$ is the temperature
1162: of the precursor averaged over the volume from $x=\xFEB$ to $x=0$,
1163: $T_1$ is the temperature at $x=\xtr$, $T_2$ is the volume-averaged
1164: temperature at $x>0$ (all temperatures are calculated from the ideal
1165: gas law~(\ref{idealgaslaw})), $\Msone$ is the sonic Mach number at $x=\xtr$,
1166: $\fcr$ is the fraction of injected thermal particles, $\pmax$ is the
1167: maximum particle momentum (i.e., the momentum at which $f(p)$ starts falling
1168: off exponentially with $p$), $\left<\gamma(x<0)\right>$ is the value of
1169: the polytropic index of particle gas, calculated from particle pressure
1170: and internal energy density as described in Appendix~\ref{compressionratio},
1171: and averaged over volume from $x=\xFEB$ to $x=0$, $\gamma_2$ is the
1172: same quantity averaged over $x>0$, $\xtr$ is the point at which the
1173: subshock starts, as defined by Eq.~(\ref{subshocklocation}),
1174: and $\xFP$ is the point defined by Eq.~(\ref{xfastpush}), at which
1175: thermal particles were introduced by the APA procedure as described
1176: in Appendix~\ref{fastpush}.
1177: }
1178: \end{deluxetable}
1179:
1180: %\clearpage
1181:
1182: \begin{deluxetable}{ lrrrrrrrr }
1183: \tabletypesize{\small}
1184: \tablecaption{Summary of Non-linear Simulation in a Hot ISM \label{sumnl_hot}}
1185: \tablewidth{0pt}
1186: \tablehead{
1187: $\heatpar$ & 0.00 & 0.10 & 0.25 & 0.50 & 0.75 & 0.95 & 1.00 & No MFA
1188: }
1189: \startdata
1190: $\rtot$ & 8.1 & 8.2 & 8.3 & 8.0 & 7.8 & 7.4 & 7.3 & 13 \\
1191: $\rsub$ & 2.92 & 2.75 & 2.55 & 2.44 & 2.22 & 2.15 & 2.12 & 2.75 \\
1192: $\Befftwo$, \muG & 62 & 60 & 55 & 44 & 32 & 17 & 14 & 21 \\
1193: $\Btrend$, \muG & 62 & 59 & 54 & 45 & 33 & 19 & 13 & - \\
1194: $\left<T(x<0)\right>$, $10^6$ K
1195: & 1.04 & 1.3 & 1.7 & 2.3 & 3.1 & 3.9 & 4.2 & 1.1 \\
1196: $T_1$, $10^6$ K & 2.0 & 6.0 & 13 & 23 & 34 & 42 & 43 & 2.7 \\
1197: $T_2$, $10^6$ K & 53 & 49 & 47 & 55 & 62 & 72 & 75 & 22 \\
1198: $\Msone$ & 10.9 & 5.8 & 3.7 & 2.6 & 2.1 & 1.9 & 1.9 & 4.7 \\
1199: $\fcr$, \% & 1.2 & 1.6 & 2.5 & 4.0 & 6.4 & 6.9 & 6.4 & 2.4 \\
1200: $\pmax / m_pc$ & 150 & 120 & 110 & 100 & 90 & 70 & 60 & 80 \\
1201: $\left<\gamma(x<0)\right>$
1202: & 1.34 & 1.34 & 1.34 & 1.34 & 1.34 & 1.34 & 1.35 & 1.34 \\
1203: $\gamma_2$ & 1.43 & 1.43 & 1.43 & 1.44 & 1.45 & 1.45 & 1.45 & 1.41 \\
1204: $\xtr / \rgzero$ & -0.04 &-0.02 & -0.02 & -0.02 &-0.03 & -0.07 & -0.05 & -0.02 \\
1205: $\xFP / \rgzero$ & -0.1 &-0.1 & -0.2 & -0.2 & -0.4 & -0.9 & -1.4 & -0.1 \\
1206: \enddata
1207: \footnotesize\tablecomments{See the note for Table~\ref{sumnl_cold} for
1208: the explanation of listed quantities.}
1209: \end{deluxetable}
1210: \clearpage
1211:
1212: Referring to Tables~\ref{sumnl_cold} and \ref{sumnl_hot}, we summarize
1213: some of the results of these models.
1214: %
1215: The effect of the turbulence dissipation into the thermal plasma is
1216: evident in the values of the pre-subshock temperature $T_1$, the
1217: downstream temperature $T_2$, and the volume-averaged precursor
1218: temperature $\left< T(x<0) \right>$ (the averaging takes place between
1219: $x=\xFEB$ and $x=0$).
1220: The temperatures were calculated from the thermal particle pressure
1221: $\Pth(x)$ using the ideal gas law equation~(\ref{idealgaslaw}).
1222: The value of $T_1$ depends drastically on the
1223: level of the turbulence dissipation $\heatpar$, increasing from
1224: $\heatpar=0$ to $\heatpar=0.5$ by a factor of 100 in the
1225: \coldismNoT case, and by a factor of 11 in the
1226: \hotismNoT case (less in the latter case, because
1227: the efficiency of the CR streaming instability in generating the
1228: magnetic turbulence is less for the smaller $M_s$ and $M_A$ for
1229: $T=10^6$~K).
1230: %
1231: The values of the temperature as high as $T_1$ are achieved upstream
1232: only near the subshock; the volume-averaged upstream temperature,
1233: $\left< T(x<0) \right>$, is significantly lower. The factors by which
1234: $\left< T(x<0) \right>$ increases in the above
1235: cases are 17 and 2.3, respectively.
1236: %
1237: The value of $\Mszero^2 / \Mazero$ is large in our models,
1238: and the estimate~(\ref{upsilonequation}) predicts that
1239: even a small amount of dissipation is enough
1240: to raise the precursor temperature significantly. This is confirmed
1241: by our results: even $\heatpar=0.1$ is enough to raise
1242: the pre-subshock temperature $T_1$ approximately 20 times in the
1243: \coldism\ case.
1244:
1245: The downstream temperature, $T_2$, varies less with changing $\heatpar$,
1246: because it is largely determined by the compression at the subshock,
1247: which is controlled by many factors as we discuss below.
1248: It is worth mentioning the case without
1249: MFA reported in Table~\ref{sumnl_hot}. Besides having a much larger
1250: compression factor than the shocks with MFA ($\rtot=13$ as opposed to
1251: $\rtot\lesssim 8$), it has a much smaller downstream temperature
1252: ($T_2=2.2\cdot 10^7$~K as opposed to $T_2 \gtrsim 5.3\cdot 10^7$~K)
1253: These effects of dissipation on the precursor temperature may be
1254: observable.
1255:
1256: %\clearpage
1257: \begin{figure}
1258: \epsscale{1.0}
1259: \plotone{f2.eps}
1260: \caption{Dependence of particle injection and magnetic field
1261: amplification on the rate of magnetic turbulence dissipation in the
1262: non-linearly modified shock with $u=5000$ \kmps\ in the \coldismNoT\
1263: and the \hotismNoT\ cases (the fully self-consistent models). The
1264: $x$-axis variable is the turbulence dissipation rate $\heatpar$
1265: (constant throughout a shock), and the plotted quantities are the
1266: amplified downstream effective magnetic field $\Befftwo$, the subshock
1267: sonic Mach number $\Msone$, the fraction of
1268: simulation particles injected into the acceleration process
1269: $\fcr$, and the ratio of amplified field to $\Btrend$ downstream
1270: from the shock.
1271: \label{fignlsum}}
1272: \end{figure}
1273: %\clearpage
1274:
1275: In Figure~\ref{fignlsum} we show results for $\fcr$, $\Msone$,
1276: $\Befftwo$, and $\Btrend$ which can be compared to the results for
1277: unmodified shocks shown in Figure~\ref{fig_unmod_inj}. For the modified
1278: shocks, the fraction of the thermal particles crossing the shock
1279: backwards for the first time, $\fcr$, clearly increases by a large
1280: factor with $\heatpar$, which can be explained by the value of $\Msone$
1281: dropping quickly below 10. One could expect that the amplified
1282: effective magnetic field $\Befftwo$ would behave similarly to the
1283: $\rtot=3.5$ case in Section~\ref{injectionvsdissipation}, i.e. that
1284: $\Befftwo$ would not decrease or even would increase for larger
1285: $\heatpar$. Instead, $\Befftwo$ behaves approximately according to the
1286: trend~(\ref{btwotrend}), as the values of $\Btrend$ from
1287: Tables~\ref{sumnl_cold} and \ref{sumnl_hot} show and the bottom panel of
1288: Fig.~\ref{fignlsum} illustrates. The important point is that, even
1289: though precursor heating causes the {\it injection efficiency} to
1290: increase substantially, the {\it efficiency of particle acceleration}
1291: and magnetic turbulence generation is hardly changed.
1292: %
1293: We base this assertion on the fact that $\Befftwo$ remains
1294: close to $\Btrend$, which was derived under the assumption that changing
1295: $\heatpar$ preserves the total energy generated by the instability, but
1296: re-distributes it between the turbulence and the thermal particles. The
1297: fact that the particle acceleration efficiency is insensitive to
1298: $\heatpar$ is directly seen in the results displayed in
1299: Figure~\ref{fig_fnp_1e4_1e6} below.
1300:
1301: Considering how much the injection rate $\fcr$ increases with
1302: $\heatpar$, and how much the upstream temperature
1303: of the thermal plasma, $T_1$, is affected by the heating,
1304: it is somewhat surprising that the
1305: trend of the amplified effective field $\Befftwo$ is unaffected. The
1306: mechanism by which the shock adjusts to the changing heating and
1307: injection in order to preserve the MFA efficiency can be understood by
1308: looking at the trend of the total compression ratio $\rtot$ and the
1309: subshock compression ratio $\rsub$ in Tables~\ref{sumnl_cold} and
1310: \ref{sumnl_hot}: they both decrease significantly for higher
1311: $\heatpar$.
1312: The decrease in $\rsub$ is easy to understand:
1313: with the turbulence dissipation operating in the precursor
1314: $\Msone$ goes down, which lowers $\rsub$. Additionally,
1315: decreasing $\Pwone$ helps to reduce $\rsub$, and
1316: with a boost
1317: of the particle injection rate, the particles returning for the first
1318: time increase in number and build up some extra pressure just upstream
1319: of the shock, which causes
1320: the flow to slow down in that region, thus reducing the ratio
1321: $\rsub$.
1322: %
1323: The decrease in $\rtot$ results from more complex processes. Here,
1324: the histories of `adolescent' particles, i.e., those particles
1325: returning upstream for the first time or for the first few times, are
1326: critical. Adolescent particles, while superthermal, are still highly
1327: anisotropic in the shock frame and how they get accelerated in the
1328: smoothed precursor just upstream of the subshock determines the number
1329: and energies of the `mature' superthermal particles, i.e., those
1330: particles with enough energy to be nearly isotropic in the shock
1331: frame. The mature particles determine the CR pressure and precursor
1332: smoothing on larger scales.
1333: %
1334:
1335: Further understanding of the shock adjustment to the changing
1336: dissipation can be gained by studying Figures \ref{fig_ubt_1e4_1e6} -
1337: \ref{fig_fnp_1e4_1e6}, in which we plot the spatial
1338: structure and the momentum-dependent quantities of the shocks in the
1339: \coldismNoT\ and the \hotismNoT\ cases for $\heatpar \in \{0; 0.5; 1\}$.
1340:
1341: %\clearpage
1342: \begin{figure}
1343: \epsscale{1.00} \plottwo{f3a.eps}{f3b.eps}
1344: \caption{ Results of non-linear simulations in the \coldismNoT\
1345: (left) and \hotismNoT\ (right) with different values of
1346: $\heatpar$. The solid, dashed and dotted lines correspond,
1347: respectively, to $\heatpar = 0$, $0.5$ and $1.0$. The plotted
1348: quantities are the bulk flow speed $u(x)$, the effective amplified
1349: magnetic field $\Beff(x)$ and the thermal gas temperature $T(x)$. The
1350: shock is located at $x=0$, and note the change from the logarithmic to
1351: the linear scale at $x=-0.05\:\rgzero$.
1352: \label{fig_ubt_1e4_1e6}}
1353: \end{figure}
1354:
1355:
1356:
1357: \begin{figure}
1358: \epsscale{1.00} \plotone{f4.eps}
1359: \caption{Enlarged subshock region in the \hotismNoT\ case. The top
1360: panel shows the flow speed normalized to $u_0$ for the
1361: $\heatpar=0.0$ and $\heatpar=0.5$ models. The middle panel shows
1362: various constituents of pressure for the $\heatpar=0.0$ run: the
1363: thermal pressure $\Pth$, the pressures of `adolescent' particles that
1364: crossed the shock 1, 2, 3 and 4 times ($P_1$, $P_2$, $P_3$ and $P_4$,
1365: respectively), the pressure of particles that have crossed 5 and more
1366: times $P_\mathrm{>5}$ and the magnetic turbulence pressure $P_w$, all
1367: of the above are normalized to the far upstream momentum flux
1368: $\Phi_\mathrm{P0}$. The bottom panel shows the same quantities for
1369: $\heatpar=0.5$.
1370: \label{fig_partpress}}
1371: \end{figure}
1372: %\clearpage
1373:
1374: Figure~\ref{fig_ubt_1e4_1e6} shows an overlap in the curves for the flow
1375: speed $u(x)$ in the $\heatpar=0$ and $\heatpar=0.5$ models, and only
1376: close to the subshock $u(x)$ falls off more rapidly towards the subshock
1377: in the $\heatpar=0.5$ case, resulting eventually in a lower
1378: $\rsub$. This means that for the high energy particles, which diffuse
1379: far upstream, the acceleration process will go on in about the same
1380: way with and without moderate turbulence dissipation
1381: (and the acceleration efficiency will be preserved with changing
1382: $\heatpar$).
1383: For lower energy particles, however, there will be observable differences in the
1384: energy spectrum. The $\heatpar=1.0$ case has a significantly smoother
1385: precursor, which is not unusual, given the lower maximal energy of the
1386: accelerated particles in this case (because of the magnetic field
1387: remaining low). The thermal gas temperatures $T(x)$ plotted in the
1388: bottom panels of Figure~\ref{fig_ubt_1e4_1e6} were calculated from the
1389: thermal pressure $\Pth(x)$ using~(\ref{idealgaslaw}) and show that the
1390: temperature becomes high well in front of the subshock.
1391:
1392: In Figure~\ref{fig_partpress} the subshock region for the \hotismNoT\
1393: case is shown enlarged, and we can compare details of the models with
1394: ($\heatpar=0.5$) and without dissipation ($\heatpar=0.0$). In the
1395: absence of turbulence dissipation, the thermal pressure $\Pth$ remains
1396: low upstream (see the middle panel), and the subshock transition is
1397: dominated by the magnetic pressure $P_w$.
1398: For $\heatpar=0.5$
1399: (the bottom panel) the thermal pressure $\Pth$ just before the shock
1400: increases enough to become comparable with the magnetic pressure, but
1401: also the heating-boosted particle injection brings up the pressures of
1402: the `adolescent' particles. As the plot shows, for $\heatpar=0.5$ the
1403: pressures produced by the first and second time returning particles
1404: ($P_1$ and $P_2$) are not small compared to $\Pth$ and $P_w$ just
1405: upstream of the shock, which contributes to the reduction of $\rsub$
1406: described above. However, the pressure of the `mature' particles,
1407: $P_\mathrm{>5}$, doesn't change much, which is a result of the
1408: non-linear response of the shock structure to the increased injection.
1409:
1410: %\clearpage
1411: \begin{figure}
1412: \epsscale{1.00}
1413: \plottwo{f5a.eps}{f5b.eps}
1414: \caption{Results of non-linear simulations in the \coldismNoT
1415: (left) and \hotismNoT\ (right) with different values
1416: of $\heatpar$. Line styles as in Fig.~\ref{fig_ubt_1e4_1e6}. The
1417: plotted quantities are: the particle distribution function in the
1418: shock frame $f(p)$ multiplied by $p^4$ (the normalization is
1419: such that $\int{4\pi p'^2 f(p') dp'}=n$, $n$ being the number density
1420: in \pcc, and $p'=p/(m_p c)$), the number of particles $n(>p)$ with
1421: momentum greater than $p$ (in \pcc) and the particle pressure (in
1422: dynes per cm$^2$) per decade of normalized momentum
1423: $dP_p/d\log_\mathrm{10}{p/(m_p c)}$. All quantities are calculated
1424: downstream at $x=+6 \rgzero$.
1425: \label{fig_fnp_1e4_1e6}}
1426: \end{figure}
1427: %\clearpage
1428:
1429:
1430: The low energy parts of the particle distribution functions shown in
1431: Figure~\ref{fig_fnp_1e4_1e6} are significantly different for models with
1432: and without dissipation in both the \coldismNoT\ and the \hotismNoT\
1433: cases. The apparent widening of the thermal peak reflects the increase
1434: in the downstream gas temperature $T_2$. The differences extend from the
1435: thermal peak to mildly superthermal momenta $0.2\: m_p c$, which shows
1436: an increased population of the `adolescent' particles with speeds up to
1437: $v\approx 0.2 c \approx 12 u_0$ when the turbulence dissipation
1438: operates. The high energy ($p>0.2 \: m_p c$) parts of the spectra for
1439: $\heatpar=0$ and $\heatpar=0.5$ are similar (except
1440: for a lower $\pmax$ due to a lower value of the amplified field in the
1441: $\heatpar=0.5$ case), confirming our
1442: assertion about the preservation of the particle acceleration
1443: efficiency. The increased population of the low-energy particles just
1444: above the thermal peak should influence the shock's X-ray emission.
1445:
1446: The characteristic concave curvature of the particle spectra above
1447: the thermal peak is clearly seen in the top panels of
1448: Figure~\ref{fig_fnp_1e4_1e6}. These shocks are strongly \NL\ and, as
1449: the bottom panels in Figure~\ref{fig_fnp_1e4_1e6} show, most of the
1450: pressure is in the highest energy particles. For these examples, 60 to
1451: 80 percent of the downstream momentum flux is in CR particles. The
1452: number of particles producing this pressure is small, however, and as
1453: the plots in the middle panels show, the fraction of particles above
1454: the thermal peak is on the order of $10^{-3}$, and the fraction of
1455: particles above 1~GeV is around $10^{-6}$ in all cases. In addition to
1456: the pressure (and energy) in the distributions shown, a sizable
1457: fraction of shock ram kinetic energy flux escapes at the FEB.
1458:
1459: To summarize,
1460: for both the unmodified (Fig.~\ref{fig_unmod_inj}) and modified
1461: (Fig.~\ref{fignlsum}) cases, $\Msone$ drops and
1462: $\fcr$ grows as $\heatpar$ increases. The surprising result is that
1463: $\Befftwo$ can increase in the unmodified shock as $\heatpar$ goes up if
1464: $\rtot$ is large enough. This indicates that the boosted injection
1465: efficiency (i.e., larger $\fcr$) outweighs the effects of field
1466: damping. This doesn't happen in the modified case (top panel of
1467: Fig.~\ref{fignlsum}) because of the \NL\ effects from the increased
1468: injection. From Fig.~\ref{fig_partpress} we see that the boosted
1469: injection results in a smoother subshock and this makes it harder for
1470: low energy adolescent particles to gain energy. Once particles reach a
1471: high enough momentum ($p\gtrsim 0.2 m_pc$; see the top panel of
1472: Fig.~\ref{fig_fnp_1e4_1e6}) they are able to diffuse far enough upstream
1473: where the boost in injection has a lesser effect.
1474: %
1475:
1476: We must emphasise again that these results are very sensitive to
1477: the physics of particle injection at the subshock. It is difficult to
1478: predict how the \NL\ results would change if a different model of
1479: injection was used, but we can refer the reader to the work of
1480: \citet{AB2006}, who performed a similar calculation using the
1481: threshold injection model with a different diffusion coefficient.
1482:
1483: The free escape boundary in our simulation was relatively close to the
1484: shock ($\xFEB = -10^5 \rgzero \approx - 3 \times 10^{-4}$ pc), and the
1485: maximum accelerated particle energy was on the order of hundreds of
1486: GeV. These quantities, chosen to save computation time, are
1487: several orders of magnitude short of typical SNR values.
1488: Nevertheless, we don't believe our results will be changed
1489: qualitatively if $\pmax$ is increased. The reason is that even with our
1490: relatively low $\pmax$, the fraction of internal energy in \rel\ particles is
1491: still large and the volume-averaged value of the
1492: polytropic index of the precursor plasma, $\left<\gamma(x<0)\right>$,
1493: as shown in Tables~\ref{sumnl_cold} and \ref{sumnl_hot}, is much closer to
1494: the $4/3$ of a fully relativistic gas than to $5/3$ for a
1495: nonrelativistic one. Increasing $\pmax$ will not lower the
1496: polytropic index of the gas any further and, consequently, the plasma
1497: compression and the subsequent acceleration efficiency will not
1498: change significantly \citep[see][]{BE99}.
1499:
1500:
1501:
1502: \section{Conclusions}
1503:
1504: \label{conclusions}
1505:
1506: We have parameterized magnetic turbulence dissipation as a fraction of
1507: turbulence energy generation and included this effect in our \MC\ model
1508: of strongly \NL\ shocks undergoing efficient DSA. The energy removed
1509: from the turbulence goes directly into the thermal particle population
1510: in the shock precursor. The \mc\ simulation \SCly\ reacts to the changes
1511: in precursor heating and adjusts the injection of thermal particles into
1512: the DSA mechanism, as well as other \NL\ effects of DSA, accordingly.
1513:
1514: Our two most important results are, first, that even a small rate
1515: ($\sim 10$\%) of turbulence dissipation can drastically increase the
1516: precursor temperature, and second, that the precursor heating boosts
1517: particle injection into DSA by a large factor. The increase in particle
1518: injection modifies the low-energy part of the particle spectrum but, due
1519: to \NL\ feedback effects, does not significantly change the overall
1520: efficiency or the high energy part of the spectrum. Both the precursor
1521: heating and modified spectral shape that occur with dissipation may have
1522: observable consequences.
1523:
1524: The fact that the shock back-reaction to the increased injection
1525: prevents the acceleration efficiency from
1526: changing significantly is a clear consequence of the non-linear
1527: structure of the system. The boosted particle injection additionally
1528: smoothes the flow speed close to the subshock, which makes it harder for
1529: particles returning upstream to gain energy. As a result, the population
1530: of the high energy particles is not much changed
1531: %
1532: (except for the decrease in the maximum particle momentum due to the
1533: reduction in the effective amplified magnetic field from dissipation)
1534: %
1535: and, because those particles carry the bulk of the CR
1536: pressure $\Pcr$ which drives the streaming
1537: instability, the amplification of the magnetic field is not
1538: strongly affected by the heating-boosted particle injection.
1539:
1540: The parameterization we use here is a simple one and a more advanced
1541: description of the turbulence damping may change our results. In our
1542: model the energy drained from the magnetic turbulence, at all
1543: wavelengths, is directly `pumped' into the thermal
1544: particles. Superthermal particles only gain extra
1545: energy due to heating because the thermal particles were more likely to
1546: return upstream and get accelerated. In a more advanced model of
1547: dissipation, where energy cascades from large-scale turbulence harmonics
1548: to the short-scale ones, the low energy CRs might gain energy directly
1549: from the dissipation. It is conceivable that cascading effects might
1550: increase the overall acceleration efficiency, the magnetic field
1551: amplification, and increase the maximum particle energy a shock can
1552: produce.
1553:
1554: It is also possible that non-resonant
1555: turbulence instabilities play an important role in magnetic
1556: field amplification \citep[e.g.,][]{PLM2006}. This opens another
1557: possibility for the turbulence dissipation to produce an increase in the
1558: magnetic field amplification. For instance, \citet{BT2005} proposed a
1559: mechanism for generating long-wavelength
1560: perturbations of magnetic fields by low energy particles. If such a
1561: mechanism is responsible for generation of a significant fraction of the
1562: turbulence that confines the highest energy particles, then the
1563: increased particle injection due to the precursor heating may raise the
1564: maximum particle energy and, possibly, the value of the amplified
1565: magnetic field.
1566:
1567: While our model is for the most part phenomenological as far
1568: as particle transport, injection and acceleration, magnetic field
1569: generation, and dissipation are concerned, it allows us to investigate
1570: the coupled nonlinear effects in a shock undergoing efficient DSA and
1571: MFA. The more efficient DSA is, the more basic considerations of
1572: momentum and energy conservation determine the shock structure and our
1573: model describes these effects fully.
1574:
1575: \acknowledgments
1576:
1577: We thank the anonymous referee for a number of very valuable comments and suggestions.
1578: %
1579: D.~C.~E. and A.~V. wish to acknowledge support from NASA grants NNH04Zss001N-LTSA
1580: and 06-ATP06-21. A.~M.~B. acknowledges support from RBRF grant 06-02-16884.
1581:
1582: % NASA ATP ( NNX07AG79G )
1583:
1584:
1585:
1586: \appendix
1587:
1588: \section{A. Comments on PIC simulations of MFA}
1589:
1590: \label{estimatesforpic}
1591:
1592: There are two basic reasons why the problem of MFA in nonlinear
1593: diffusive shock acceleration (NL-DSA) is particularly difficult for
1594: particle-in-cell (PIC) simulations. The
1595: first is that PIC simulations must be done fully in three dimensions to
1596: properly account for cross-field diffusion. As \citet{JJB98} proved from
1597: first principles, PIC simulations with one or more ignorable dimensions
1598: unphysically prevent particles from crossing magnetic field lines. In
1599: all but strictly parallel shock geometry,\footnote{Parallel geometry is
1600: where the upstream magnetic field is parallel to the shock normal.} a
1601: condition which never occurs in strong turbulence, cross-field
1602: scattering is expected to contribute importantly to particle injection
1603: and must be fully accounted for if injection from the thermal background
1604: is to be modeled accurately.
1605:
1606: The second reason is that, in \nonrel\ shocks, NL-DSA spans large
1607: spatial, temporal, and momentum scales. The range of scales is more
1608: important than might be expected because DSA is intrinsically efficient
1609: and \NL\ effects tend to place a large fraction of the particle pressure
1610: in the highest energy particles (see
1611: Fig.~\ref{fig_fnp_1e4_1e6}). The highest energy particles, with the
1612: largest diffusion lengths and longest acceleration times, feedback on
1613: the injection of the lowest energy particles with the shortest scales.
1614: The accelerated particles exchange their momentum and energy with the
1615: incoming thermal plasma through the magnetic fluctuations coupled to the
1616: flow. This results in the flow being decelerated and the plasma being
1617: heated. The structure of the shock, including the subshock where fresh
1618: particles are injected, depends critically on the highest energy
1619: particles in the system.
1620:
1621: A plasma simulation must resolve the electron skin depth,
1622: $c/\ElPlasmafreq$, i.e., $\Lcell < c/\ElPlasmafreq$, where
1623: $\ElPlasmafreq=[4 \pi n_e e^2/m_e]^{1/2}$ is the electron plasma
1624: frequency and $\Lcell$ is the simulation cell size. Here, $n_e$ is
1625: the electron number density, $m_e$ is the electron mass and $c$ and
1626: $e$ have their usual meanings. The simulation must
1627: also have a time step small compared to $\ElPlasmafreq^{-1}$, i.e.,
1628: $\tstep < \ElPlasmafreq^{-1}$.
1629: %
1630: If we wish to follow the acceleration of protons in DSA to the TeV
1631: energies present in SNRs we must have a simulation box that is as large
1632: as the upstream diffusion length of the highest energy protons, i.e.,
1633: $\kappa(\mathrm{\Emax})/u_0 \sim r_g(\Emax) c / (3 u_0)$, where $\kappa$
1634: is the diffusion coefficient, $r_g(\Emax)$ is the gyroradius of a \rel\
1635: proton with the energy $\Emax$, $u_0$ is the shock speed, and we have
1636: assumed Bohm diffusion.
1637: %
1638: The simulation must also be able to run for as long as the acceleration
1639: time of the highest energy protons, $\tacc(\Emax) \sim \Emax c /(eBu_0^2)$.
1640: %
1641: Here, $B$ is some average magnetic field.
1642: %
1643: The spatial condition gives
1644: %
1645: %
1646: \begin{equation}
1647: \frac{\kappa(E_\mathrm{max})/u_0}{(c/\ElPlasmafreq)} \sim 6\xx{11}
1648: \left ( \frac{\Emax}{\mathrm{TeV}} \right )
1649: \left ( \frac{u_0}{1000 \, \mathrm{km \, s}^{-1}} \right )^{-1}
1650: \left ( \frac{B}{\mu \mathrm{G}} \right )^{-1}
1651: \left ( \frac{n_e}{\mathrm{cm}^{-3}}\right )^{1/2}
1652: \left ( \frac{f}{1836}\right )^{1/2}
1653: \ ,
1654: \end{equation}
1655: %
1656: for the number of cells {\it in one dimension}. The factor $f=m_p/m_e$
1657: is the proton to electron mass ratio. From the acceleration
1658: time condition, the required number of time steps is,
1659: %
1660: \begin{equation}
1661: \frac{\tacc(E_\mathrm{max})}{\ElPlasmafreq^{-1}} \sim 6\xx{14}
1662: \left ( \frac{\Emax}{\mathrm{TeV}} \right )
1663: \left ( \frac{u_0}{1000\,\mathrm{km\,s}^{-1}}\right )^{-2}
1664: \left ( \frac{B}{\mu \mathrm{G}} \right )^{-1}
1665: \left ( \frac{n_e}{\mathrm{cm}^{-3}}\right )^{1/2}
1666: \left ( \frac{f}{1836}\right )^{1/2}
1667: \ .
1668: \end{equation}
1669: %
1670: Even with $f=1$ these numbers are obviously far beyond any conceivable
1671: computing capabilities and they show that approximate methods are
1672: essential for studying NL-DSA.
1673:
1674: One approximation that is often used is a hybrid PIC simulation where
1675: the electrons are treated as a background fluid. To get the
1676: estimate of the requirements in this case we can take the
1677: minimum cell size as the thermal proton gyroradius,
1678: $\rgzero = c \sqrt{2 m_p \Eth}/(e B)$.
1679: Now, the number of cells, {\it again in one dimension,} is:
1680: \begin{equation}
1681: \label{eq:cell}
1682: \frac{\kappa(\Emax)/u_0}{\rgzero} \sim 7\xx{7}
1683: \left ( \frac{\Emax}{\mathrm{TeV}} \right )
1684: \left ( \frac{u_0}{1000 \, \mathrm{km \, s}^{-1}} \right)^{-1}
1685: \left ( \frac{\Eth}{\mathrm{keV}} \right )^{-1/2}
1686: \ .
1687: \end{equation}
1688: %
1689: The time step size must be $\tstep < \protongyrofreq^{-1}$,
1690: where $\protongyrofreq=eB/m_{p}c$ is the thermal proton gyrofrequency.
1691: This gives the number of time steps to reach 1 TeV,
1692: %
1693: \begin{equation}
1694: \label{eq:step}
1695: \frac{\tacc(\Emax)}{\protongyrofreq^{-1}} \sim 1\xx{8}
1696: \left ( \frac{\Emax}{\mathrm{TeV}} \right )
1697: \left ( \frac{u_0}{1000\,\mathrm{km\,s}^{-1}}\right )^{-2}
1698: \ .
1699: \end{equation}
1700: %
1701: These combined spatial and temporal requirements, even for the most
1702: optimistic case of a hybrid simulation with an unrealistically large
1703: $\tstep$, are well beyond existing computing capabilities unless
1704: %exceptionally clever shortcuts are devised or
1705: a maximum energy well below 1 TeV is used.
1706:
1707: Since the three-dimensional requirement is fundamental and relaxing it
1708: eliminates cross-field diffusion, restricting the energy range is the
1709: best way to make the problem assessable to hybrid PIC
1710: simulations. However, since producing \rel\ particles from \nonrel\ ones
1711: is an essential part of the NL problem, the energy range must
1712: comfortably span $m_p c^2$ to be realistic. If $\Emax=10$\,GeV is used,
1713: with $u_0=5000$\,\kmps, and $\Eth= 10$\,MeV, equation~(\ref{eq:cell})
1714: gives $\sim 1400$ and equation~(\ref{eq:step}) gives $\sim 4\xx{4}$.
1715: Now, the computation may be possible, even with the 3-D requirement, but
1716: the hybrid simulation can't fully investigate MFA since electron return
1717: currents are not modeled. The exact microscopic description of the
1718: system is not currently feasible.
1719:
1720: It's hard to make a comparison in run-time between PIC simulations
1721: and the \MC\ technique used here since we are not aware of any published
1722: results of 3-D PIC simulations of \nonrel\ shocks that follow particles
1723: from fully \nonrel\ to fully \rel\ energies. A direct comparison of 1-D
1724: hybrid and \MC\ codes was given in \citet{EGBS93} for energies
1725: consistent with the acceleration of diffuse ions at the quasi-parallel
1726: Earth bow shock. Three-dimensional hybrid PIC results for \nonrel\
1727: shocks were presented in \citet{GE2000} and these were barely able to
1728: show injection and acceleration given the computational limits at that
1729: time. As for the \MC\ technique, all of the 16 nonlinear simulations
1730: presented in this paper were completed in approximately 4 days on a
1731: parallel computing cluster, employing around 30 processors; enough
1732: statistical information was accumulated to restrict the uncertainty in
1733: the self-consistent value of $\rtot$ to about 5 percent.
1734: %
1735: Increasing the
1736: dynamic range of the simulations to SNR-like energies ($\Emax
1737: \approx 1$\,PeV) would require 2-4 times more computation time.
1738: %
1739: Thus, realistic \mc\ SNR models are possible with modest
1740: computing resouces.
1741:
1742:
1743: Despite these limitations, PIC simulations are the only way of \SCly\
1744: modeling the plasma physics of collisionless shocks.
1745: %
1746: In particular, the injection of thermal particles in the large amplitude
1747: waves and time varying structure of the subshock can only only be
1748: determined with PIC simulations
1749: \citep[e.g.,][]{NPS2008,Spitkovsky2008}. Injection is one of the most
1750: important aspects of DSA and one
1751: where analytic and \mc\ techniques have large uncertainties.
1752:
1753:
1754: \section{B. Analytic Precursor Approximation procedure}
1755:
1756: \label{fastpush}
1757:
1758: Here we describe our \FP\ (APA) where we introduce
1759: thermal particles into the MC simulation,
1760: not far upstream, but at some position $\xFP<0$ close
1761: to the subshock. This
1762: procedure has two purposes. First, it saves computational
1763: time because we don't have to trace these particles
1764: along the extended shock precursor. Second, we use the APA
1765: to simulate the turbulence dissipation in the shock precursor:
1766: with this procedure we incorporate the analytic description
1767: of the heating of the thermal gas due to the turbulence dissipation
1768: into the Monte Carlo model of particle transport.
1769:
1770: In the next two subsections we
1771: describe how the momentum and energy fluxes in shock precursor are
1772: calculated in the absence of turbulence dissipation (part~\ref{noapa}),
1773: and then explain how this scheme is changed when the \FP\ procedure is
1774: invoked to model the thermal plasma heating by turbulence dissipation
1775: (part~\ref{yesapa}).
1776:
1777:
1778: \subsection{B.1. Precursor without Analytic Approximation}
1779:
1780: \label{noapa}
1781:
1782: \subsubsection{B.1.1. Inherent Quasi-Adiabatic Heating
1783: and Subshock Definition in Monte Carlo Simulation}
1784:
1785: \label{mcadiabatic}
1786:
1787: It is worth pointing out that even if the turbulence dissipation
1788: is not included in our simulation, particles in the precursor
1789: will still be weakly heated.
1790: %
1791: Just like an ideal collisionless gas put in a slowly shrinking volume
1792: will heat up adiabatically due to elastic collisions of particles with
1793: in-moving walls, the particles in our MC simulation traveling in a
1794: compressing shock precursor will heat up according to the adiabatic
1795: law $\Pth \propto \rho^{\gamma} \propto u^{-\gamma}$ due to
1796: elastic scattering in the decelerating local plasma frame.
1797: This will be true as long as the particles have enough time to adjust
1798: to the changing flow speed. Quantitatively, the criterion
1799: for adiabatic heating is
1800: \begin{equation}
1801: \label{adiabaticcondition}
1802: \taucoll \ll \taucomp,
1803: \end{equation}
1804: where the collision time, $\taucoll$, is the ratio of the mean
1805: free path to the particle speed, (for nonrelativistic particles
1806: $\taucoll=m_p c/(e \Beff)$),
1807: %
1808: and the compression time $\taucomp$ is the temporal scale on which the
1809: speed of plasma that the particle is traveling in changes significantly,
1810: that is, by a value comparable to the particle speed $v$. Assuming $v
1811: \ll u$, $\taucomp$ is approximately the ratio of the length on which the
1812: flow speed changes by $v$ to the advection speed $u$:
1813: $\taucomp=(v/|(du/dx)|) / u$. It also follows from the definition of
1814: $\taucoll$ that the angular distribution of the particles remains
1815: isotropic if (\ref{adiabaticcondition}) is true. The
1816: condition~(\ref{adiabaticcondition}) is therefore
1817: %
1818: \begin{equation}
1819: \left| \frac{du}{dx} \right| \ll \frac{v e \Beff}{u m_p c},
1820: \end{equation}
1821: %
1822: restricting the flow speed gradient to small enough values, where
1823: adiabatic compression operates.
1824:
1825: If the magnitude of the flow speed gradient $|du/dx|$ is too large
1826: and (\ref{adiabaticcondition}) doesn't hold, the increase
1827: of particle energies will be faster than adiabatic, and the angular
1828: distribution function will become non-isotropic, with more particles
1829: moving against the gradient than along it in the plasma frame. One
1830: obvious place where this occurs is the subshock.
1831: %
1832: In our simulation, the subshock in the non-linear,
1833: self consistent solution gains a finite width, but the flow speed jump
1834: at the subshock remains strong and rapid enough to efficiently
1835: accelerate particles. We define the point at which the
1836: transition from the adiabatic to the non-adiabatic regime occurs,
1837: $\xtr$, by the following condition:
1838: %
1839: \begin{equation}
1840: \taucoll(\xtr) = \frac13 \taucomp(\xtr),
1841: \end{equation}
1842: %
1843: or, in terms of quantities available in the simulation,
1844: %
1845: \begin{equation}
1846: \label{subshocklocation}
1847: \left. \frac{du}{dx} \right|_{\xtr} =
1848: - \frac13 \frac{v e \Beff(\xtr)}{u(\xtr) m_p c}
1849: ,
1850: \end{equation}
1851: %
1852: where $v=\sqrt{2 k_B T(\xtr)/m_p}$,
1853: in which case $|\xtr|$ is comparable to the local
1854: convective mean free path of
1855: a thermal particle. The factor $1/3$ is chosen arbitrarily but our
1856: results are insensitive to it. Close to the shock, at $\xtr < x < 0$,
1857: the flow speed drops so rapidly that the particles are heated in a
1858: non-adiabatic fashion. We can think of the non-adiabatic region as
1859: a subshock with a finite thickness $|\xtr|$, and it is then
1860: reasonable to define the pre-subshock quantities (denoted with index
1861: `1') as the values at $x=\xtr$: $u_1 = u(\xtr)$, $\rsub =
1862: u(\xtr)/u(x>0)$, etc.
1863: %
1864: We note that Equation~\ref{subshocklocation} is only used to define
1865: the position of the subshock an is not used in any calculations.
1866:
1867: \subsubsection{B.1.2. Direct Calculation of Momentum and Energy Fluxes}
1868:
1869: \label{calculatingfluxes}
1870:
1871: The flux of momentum $\momentumflux(x)$ defined in~(\ref{mfasmoment}) is
1872: used in our simulation to calculate the smoothing of the precursor
1873: plasma flow $u(x)$, and the flux of energy $\energyflux(x)$ defined
1874: in~(\ref{efasmoment}) is used to calculate the compression ratio $\rtot$
1875: consistent with the particle acceleration. The moments of the particle
1876: distribution function in~(\ref{mfasmoment}) and (\ref{efasmoment}) are
1877: the components of the stress tensor. If the plasma heating by turbulence
1878: dissipation is not modeled, and the \FP\ procedure is not
1879: performed, then these moments are calculated in our simulation as
1880: described below. At select positions on the numerical grid spanning
1881: from far upstream to some position downstream from the subshock
1882: we sum the contributions of the particles that cross these positions to
1883: calculate the following:
1884: \begin{eqnarray}
1885: \label{mfcalc}
1886: \int p_x v_x f(x,\pvector)d^3p &=&
1887: \sum\limits_{i} p_\mathrm{i,\,x} v_\mathrm{i,\,x} w_i,\\
1888: \label{efcalc}
1889: \int K v_x f(x,\pvector)d^3p &=&
1890: \sum\limits_{i} K_i v_\mathrm{i,\,x} w_i.
1891: \end{eqnarray}
1892: Here the summation is taken over all particles crossing the position $x$
1893: at which the moments are calculated. The index $i$ represents the
1894: considered particle, $p_x$ ($v_x$) is the $x$-component of a
1895: particle's of momentum (velocity),
1896: $K$ is the kinetic energy, all measured in the shock frame, and $w$ is
1897: the weight of the particle defined as
1898: \begin{equation}
1899: w_i = \left| \frac{u_0}{v_\mathrm{i,\,x}} \right| \frac{n_0}{N_p}.
1900: \end{equation}
1901: In this definition the ratio $|u_0/{v_\mathrm{i,\,x}}|$ is the weighting
1902: factor accounting for the fact that in our simulation particles crossing
1903: the position $x$ at some angle to the flow do it less frequently than
1904: particles crossing parallel to the flow, $n_0$ is the upstream number
1905: density of the plasma and $N_p$ is the number of simulation particles.
1906:
1907: If the particle distribution is isotropic in the local plasma
1908: frame moving at speed $u(x)$ relative to the shock, then the quantities
1909: calculated in (\ref{mfcalc}) and (\ref{efcalc}) can be expressed in the
1910: following way:
1911: \begin{eqnarray}
1912: \label{momentumviapressures}
1913: \int p_x v_x f(x,\pvector)d^3p &=&
1914: \rho(x) u^2(x) + P_p(x), \\
1915: \label{energyviapressures}
1916: \int K v_x f(x,\pvector)d^3p &=&
1917: \frac{1}{2} \rho(x) u^3(x) + w_p(x)u(x) \ ,
1918: \end{eqnarray}
1919: where $P_p(x)$ is the pressure and $w_p(x)$ is the enthalpy
1920: of the particles. In the
1921: vicinity of the subshock, however, the isotropy
1922: assumption breaks down (see the discussion in Appendix~\ref{mcadiabatic}
1923: and the note on anisotropy in Appendix~\ref{special_th_vs_cr}), and the
1924: concept of isotropic pressure is not applicable. The fact that we
1925: directly calculate the moment of the distribution function
1926: in~(\ref{mfasmoment}) by evaluating the sum~(\ref{mfcalc}), instead of
1927: approximating the momentum flux with~(\ref{momentumviapressures}),
1928: ensures that we properly account for the effects of the anisotropy of
1929: particle distribution. This turns out to be important for
1930: self-consistently determining the flow speed $u(x)$ near the subshock,
1931: which controls the subshock compression and the subsequent particle
1932: injection and acceleration efficiency.
1933:
1934: When plasma heating by turbulence dissipation is modeled, we replace the
1935: calculation~(\ref{mfasmoment}) with an analytic approximation assuming
1936: isotropic particle pressure, but we take special care to be certain that
1937: this approximation is only done far enough from the shock, where the
1938: isotropy approximation is applicable. This procedure is described below.
1939:
1940:
1941:
1942: \subsection{B.2. Modeling Precursor Heating with the \FPshort}
1943:
1944: \label{yesapa}
1945:
1946: \subsubsection{B.2.1. Particle Introduction Position}
1947:
1948: The position at which the thermal particles are introduced must be close
1949: enough to the shock so that the analytic description
1950: of heating applies to most of the precursor extent, but far enough away
1951: from the non-adiabatic region $\xtr < x < 0$, so that the analytic
1952: approximation remains valid where applied. In our simulation we chose
1953: the particle introduction position, $\xFP$, so that the
1954: condition~(\ref{adiabaticcondition}) is only marginally valid at this
1955: position. We formalize it as
1956: \begin{equation}
1957: \taucoll(\xFP) = M \taucomp(\xFP),
1958: \end{equation}
1959: which is equivalent to
1960: \begin{equation}
1961: \label{xfastpush}
1962: \left. \frac{du}{dx} \right|_{\xFP} = - M \frac{v e \Beff(\xFP)}{u(\xFP) m_p c}.
1963: \end{equation}
1964: where we chose $M = 0.1$ and $v = \sqrt{2 k_B T_0/m_p}$. At $x<\xFP$ we
1965: describe the thermal particle distribution function as a
1966: Maxwellian with the temperature defined by (\ref{pressuregrowth}) and
1967: (\ref{idealgaslaw}), and at $x>\xFP$ we use the Monte Carlo
1968: simulation to describe the more complex particle dynamics.
1969:
1970:
1971: \subsubsection{B.2.2. Momentum Space Distribution of Introduced Particles}
1972:
1973: In order to include the effects of heating in the model, we must
1974: introduce thermal particles at $\xFP$ as if they were heated in the
1975: precursor, i.e., their temperature $T(\xFP)$ must be determined
1976: by~(\ref{pressuregrowth}) and (\ref{idealgaslaw}). We therefore choose
1977: the magnitude of every particle's momentum $p$ in the local plasma frame
1978: distributed according to Maxwell-Boltzmann distribution with probability
1979: density
1980: \begin{equation}
1981: f(p) = \frac{4}{\sqrt{\pi}} \left( \frac{1}{2 m k_B T(\xFP)} \right)^{3/2}
1982: p^2 \exp{\left({-\frac{p^2}{2 m k_B T(\xFP)}}\right)}.
1983: \end{equation}
1984:
1985: The angular distribution of momenta of the
1986: introduced particles is a major issue of concern in doing simulation
1987: like ours because it determines the particle
1988: injection rate. We are replacing the dynamics of particles at $x<\xFP$
1989: with an analytical description, consequently we must distribute
1990: particles in $p$-space at $\xFP$ the way they would be distributed
1991: having traveled from far upstream and reaching $\xFP$ {\it for the first
1992: time}. This is equivalent to calculating a $p$-space distribution of
1993: particles incident on a {\it fully absorbing boundary} at $\xFP$ after
1994: scattering in a non-uniform flow $u(x)$. This is easy to do analytically
1995: if all particles have a plasma frame speed $v$ less than the flow speed
1996: $u(\xFP)$ (because then all particles crossing position $\xFP$ do it for
1997: the first and the last time), and fairly complicated otherwise. We
1998: assume $v<u(\xFP)$ in further reasoning, which is justified by the fact
1999: that we find $M_s(\xFP) \gtrsim 3$ in most cases.
2000:
2001: As was stated earlier, we assume that the angular distribution of
2002: momenta of the introduced thermal particles is isotropic in the plasma
2003: frame. When these particles cross a position fixed in the shock
2004: frame, their flux must be `weighted' to account for the fact that the
2005: number of particles arriving at $\xFP$ in a unit time is proportional to
2006: the cosine of the angle that their shock frame velocity
2007: $\vvector_\mathrm{sf}$ makes with the $x$-axis.
2008: This can be done by assuming a probability density of
2009: $\mu = v_x/v$ ($v$ is the magnitude of the particle plasma frame
2010: velocity and $v_x$ its $x$-component) as
2011: \begin{equation}
2012: \label{fluxweighting}
2013: f(\mu)=\frac12 \left(1 + \mu \frac{v}{u}\right),
2014: \end{equation}
2015: where $u=u(\xFP)$. It is normalized so that the probability
2016: $Pr(\mu_0 < \mu < \mu_0 + d\mu_0) = f(\mu_0) d\mu_0$, and the functional
2017: form of (\ref{fluxweighting}) comes from the assumption that $f(\mu)
2018: \propto v_\mathrm{sf,\:x}=u + \mu v$.
2019:
2020:
2021:
2022: \subsubsection{B.2.3. Heating of the Upstream Plasma}
2023:
2024: After the thermal particles are introduced at $\xFP$ and start to
2025: propagate in the shocked flow, we have to
2026: calculate the momentum and energy fluxes throughout the shock, for use
2027: in our iterative procedure. Because we didn't propagate the thermal
2028: particles at $x<\xFP$, and because we must model the momentum flux
2029: redistribution between the turbulence and the thermal particles due to
2030: heating, we calculate the corresponding moments of particle
2031: distribution function the following way:
2032: \begin{eqnarray}
2033: \label{apamomentum}
2034: \int p_x v_x f(x,\pvector)d^3p &=&
2035: \left\{
2036: \begin{array}{ll}
2037: \sum\limits_{all\:i} p_\mathrm{i,\,x} v_\mathrm{i,\,x} w_i, & \mathrm{if} \; x > \xFP, \\
2038: \rho(x) u^2(x) + \Pth(x) +
2039: \sum\limits_{i \in CR} p_\mathrm{i,\,x} v_\mathrm{i,\,x} w_i, & \mathrm{if} \; x < \xFP,
2040: \end{array}
2041: \right. \\
2042: \label{apaenergy}
2043: \int K v_x f(x,\pvector)d^3p &=&
2044: \left\{
2045: \begin{array}{ll}
2046: \sum\limits_{all\:i} K_i v_\mathrm{i,\,x} w_i, & \mathrm{if} \; x > \xFP, \\
2047: \displaystyle\frac12 \rho(x) u^3(x) + \displaystyle\frac{\gamma}{\gamma-1}\Pth(x)u(x) +
2048: \quad & \\ \qquad \qquad \qquad \qquad \qquad
2049: + \sum\limits_{i \in CR} K_i v_\mathrm{i,\,x} w_i, & \mathrm{if} \; x < \xFP.
2050: \end{array}
2051: \right.
2052: \end{eqnarray}
2053: The summation for $x>\xFP$ is taken over all the particles crossing position
2054: $x$, and for $x<\xFP$ the summation index $i\in CR$ only includes the CR (i.e., injected)
2055: particles, while the contribution of the thermal particles is replaced by the
2056: analytic approximation of this contribution. The thermal pressure in this approximation
2057: is taken from the solution of~(\ref{pressuregrowth}).
2058: For $\heatpar=0$ the equations~(\ref{apamomentum})
2059: and (\ref{apaenergy}) produce the same results
2060: (within intrinsic random deviations of the Monte Carlo code)
2061: as the calculation~(\ref{mfcalc}) and (\ref{efcalc}).
2062:
2063: In the region $\xFP < x < 0$ the heating due to $\Lbar$ is ignored.
2064: This may, in principle,
2065: underestimate the heating of the upstream gas, but we find that this
2066: is a negligible effect. We prove it by running a simulation with
2067: another $\xFP$, even farther away from the subshock, and making sure
2068: that the results are not affected significantly.
2069:
2070:
2071:
2072:
2073: \section{C. Note on Differentiation between Thermal and CR particles}
2074:
2075: \label{special_th_vs_cr}
2076:
2077: As was mentioned earlier, we call a particle a thermal particle or a CR
2078: one based on its history: a CR particle is one that has gotten injected
2079: into the acceleration process by having crossed the shock from the
2080: downstream to the upstream region at least once\footnote{Although
2081: the shock gains a finite width in the Monte Carlo simulation, we
2082: define the backward shock crossing as moving from $x>0$ to $x<0$.}.
2083: This criterion is used
2084: at two places in the model. First, it is used to calculate the spectrum
2085: of pressure driving the CR streaming instability $\Pcrhat(x,p)$: only
2086: the contribution of injected particles is included in
2087: $\Pcrhat(x,p)$. Second, when we calculate the thermal particle
2088: pressure $\Pth(x)$ at $x<\xFP$ using (\ref{dampingparameterization})
2089: and (\ref{pressuregrowth}), and then introduce thermally distributed
2090: particles at $x=\xFP$ and continue the calculation
2091: of pressure for $x>\xFP$ from their trajectories as they elastically
2092: scatter in the flow, we implicitly
2093: assume that the dissipated energy of the turbulence goes directly
2094: into the thermal energy of the particles that have not been injected,
2095: i.e., thermal particles in our definition.
2096:
2097:
2098: If a description of particle-wave interactions in strong turbulence
2099: existed that explicitly described how particles of different energies
2100: participated differently in the instability generation and the
2101: turbulence dissipation, a criterion like ours would not be
2102: needed. However, such a description is not available and we believe that our way
2103: of separating thermal and superthermal (CR) particles for purposes of
2104: describing the instability growth and the turbulence dissipation grasps
2105: the essential non-linear effects in the shock structure.
2106: PIC simulations are, in principle, able to tackle this problem
2107: exactly but, as we mentioned earlier, they are extremely computationally
2108: expensive.
2109:
2110: We would like to point out an important consequence of our using the
2111: `thermal leakage' model of particle injection into
2112: the acceleration process. Our simulation follows histories of charged
2113: particles from their `childhood', when their speeds in the plasma frame
2114: are small compared to the bulk flow speed, to `maturity', when they
2115: become relativistic. Unlike most semi-analytic descriptions of DSA, our
2116: model doesn't skip the `adolescence' stage of particles, when after one
2117: or a few shock crossings the particles have speeds comparable to or
2118: slightly greater than the bulk flow speed. In the absence of these
2119: `adolescent' particles, it is typically assumed that the jump in only
2120: the thermal particle pressure across the subshock determines the
2121: strength of the latter, and the superthermal part of the particle
2122: spectrum is continuous at the subshock and does not influence it. But
2123: the `adolescent' particles that the Monte Carlo model does describe are
2124: not energetic enough to be insensitive to the subshock, and at the same
2125: time they have a strong anisotropy in the plasma frame, and therefore do
2126: not obey the Rankine-Hugoniot relations. This modifies the conservation
2127: laws across the subshock, because in the total kinetic pressure
2128: $P_p(x)=\Pth(x) + \Pcr(x)$ the term $\Pcr(x)$ is not continuous at the
2129: subshock due to the contribution of the intermediate energy particles
2130: that it contains.
2131:
2132:
2133:
2134:
2135:
2136:
2137:
2138: \section{D. Compression ratio, Turbulence and Escaping Particles}
2139:
2140: \label{compressionratio}
2141:
2142: Equation (10) in \citet{EMP90} relates the fraction of energy flux
2143: carried away by escaping particles $\qesc$ to the total shock
2144: compression ratio, $\rtot$. It assumes no
2145: magnetic field amplification and a polytropic index of downstream gas
2146: equal to $5/3$ (in other words, neglects the effect of relativistic
2147: particles on the overall compressibility of the gas). In order to search
2148: for a $\rtot$ consistent with the shock structure and with
2149: particle escape, we derive a similar relationship that
2150: would account for the presence of magnetic turbulence and for the
2151: contribution of the relativistic particles. The problem is complicated
2152: by having particles of intermediate (mildly relativistic) energies and
2153: by the value of the magnetic field dependent on the particle
2154: acceleration.
2155:
2156: Writing equations (\ref{fluxmass}), (\ref{fluxmomentum}) and
2157: (\ref{fluxenergy}) for a point downstream of the shock, sufficiently
2158: far from it, so that the distribution of particle momenta is isotropic,
2159: and the approximations~(\ref{momentumviapressures}) and
2160: (\ref{energyviapressures}) are valid, and
2161: denoting the corresponding quantities by index `2', we get:
2162: \begin{eqnarray}
2163: %
2164: \label{cons1}
2165: \rho_2 u_2 &=& \rho_0 u_0, \\
2166: %
2167: \label{cons2}
2168: \rho_2 u_2^2 + \Pptwo + \Pwtwo &=& \rho_0 u_0^2 + \Ppzero + \Pwzero\equiv\upstreammomentumflux, \\
2169: %
2170: \label{cons3}
2171: \frac12 \rho_2 u_2^3 + \wptwo u_2 + \Fwtwo + \Qesc&=&
2172: \frac12 \rho_0 u_0^3 + \wpzero u_0 + \Fwzero\equiv\upstreamenergyflux.
2173: \end{eqnarray}
2174: %
2175: The particle gas enthalpy $w_p$ is $w_p=\epsilon_p + P_p$, and the
2176: internal energy $\epsilon_p$ of gas is proportional to
2177: the pressure $P_p$. Introducing the quantity $\gammabar$ so that
2178: $\epsilon_p = P_p/(\gammabar-1)$, one can write
2179: \begin{equation}
2180: \label{fppp}
2181: w_p u=\frac{\gammabar}{\gammabar-1}P_p u
2182: \end{equation}
2183: %
2184: The value of $\gammabar$ is averaged over the whole particle
2185: spectrum, and it ranges between $5/3$ for a nonrelativistic
2186: and $4/3$ for an ultra-relativistic gas.
2187: The local value of $\gammabar$ can be easily calculated in our
2188: code from the particle distribution, along with
2189: $P_p$ and $\epsilon_p$, as $\gammabar=1 + P_p / \epsilon_p$.
2190: %
2191: Similarly, one can define $\deltabar=F_w/(u P_w)$ and calculate a local
2192: value of $\deltabar$ anywhere in the code in order to express
2193: \begin{equation}
2194: \label{fwpw}
2195: F_w=\deltabar \cdot P_w u \ .
2196: \end{equation}
2197: %
2198: For $V_G \ll u$ and \Alf ic turbulence, one expects $\deltabar
2199: \approx 3$, [see eq.~(\ref{fwdef})].
2200:
2201: Substituting (\ref{fppp}) and (\ref{fwpw}) into the above equations
2202: and introducing $\rtot=u_0 / u_2$, we can eliminate $\rho_2$ using
2203: (\ref{cons1}) and $\Pptwo$ using (\ref{cons2}), which allows us to
2204: express from (\ref{cons3}) the quantity $\qesc \equiv \Qesc /
2205: \upstreamenergyflux$ as
2206: \begin{equation}
2207: \label{qescrtot}
2208: \qesc = 1 + \frac{A/\rtot^2 - B/\rtot}{C},
2209: \end{equation}
2210: where
2211: \begin{eqnarray}
2212: A&=& \frac{\gammabar_2 + 1}{\gammabar_2 - 1} ,\\
2213: B&=& \frac{2 \gammabar_2}{\gammabar_2 - 1}
2214: \left( 1 + \frac{\Ppzero + \Pwzero - \Pwtwo}{\rho_0 u_0^2}\right) +
2215: \frac{2 \deltabar_2 \Pwtwo}{ \rho_0 u_0^2 },\\
2216: C&=& 1 + \frac{2 \gammabar_0}{\gammabar_0 - 1}\frac{\Ppzero}{\rho_0 u_0^2} +
2217: \frac{2 \deltabar_0 \Pwzero}{\rho_0 u_0^2}.
2218: \end{eqnarray}
2219: Note that $\rho_0 u_0^2 / \Ppzero = \gammabar_0 M_s^2$, where
2220: $\gammabar_0=\gamma=5/3$ due to the absence of CRs far upstream,
2221: and, because we assume a seed turbulence far upstream, that
2222: provides a Bohm regime of scattering to all particles,
2223: one can write that far upstream $\Delta B \approx B_0$, making
2224: $\rho_0 u_0^2 / \Pwzero=2 M_A^2$.
2225:
2226: The quantity $\qesc$ is readily available in the simulation after
2227: the end of any iteration.
2228: Comparing it to the value predicted by (\ref{qescrtot}), we
2229: evaluate the self-consistency of the solution and make the correction to
2230: $\rtot$, if necessary, for further iterations. For making these
2231: corrections it is helpful to use in the simulation the inverse of
2232: (\ref{qescrtot}), the physically relevant branch of which is
2233: \begin{equation}
2234: \label{rtotqesc}
2235: \rtot = \frac{2 A}{B - \sqrt{B^2 - 4 A C (1-\qesc)}}.
2236: \end{equation}
2237:
2238: It is important to emphasise here that an iterative procedure
2239: similar to~(\ref{iteration}) is still
2240: required to find the compression ratio $\rtot$ of a non-linearly
2241: modified shock, because quantities $\qesc$, $\Pwtwo$ and $\gammabar_2$
2242: depend on $\rtot$, so (\ref{rtotqesc}) only provides a practical
2243: way to perform the iterations.
2244:
2245:
2246:
2247:
2248:
2249:
2250: \section{E. Note on Subshock Properties in the Presence of MFA}
2251:
2252: \citet{KJG2002} showed, using a thermal leakage injection model
2253: similar to what is used here, that in a non-linearly modified shock with
2254: Bohm diffusion and a sonic Mach number $\Mszero$, the subshock sonic
2255: Mach number scales as $\Msone \sim 2.9 \Mszero^{0.13}$ with the
2256: corresponding subshock compression ratio $\rsub =
2257: 4/(1+3/\Msone^2)$. Numerically, for $\Mszero=400$ it gives $\Msone
2258: \approx 6.3$ and $\rsub \approx 3.7$.
2259:
2260: The model of \citet{KJG2002} does not include magnetic field
2261: amplification and, as we have shown, in
2262: our case the values of the subshock sonic Mach number and compression
2263: ratio are quite different. For instance, for our model with MFA, but
2264: without magnetic turbulence damping ($\heatpar=0$) we have found that
2265: $\Mszero = 426$ results in $\Msone \approx 44$
2266: and $\rsub \approx 3$. The
2267: value of $\Msone$ in the absence of precursor plasma heating is
2268: determined by the high CR pressure $\Pcr(x)$, and the value of $\rsub$
2269: in the high Mach number shock with MFA is largely controlled by the
2270: pressure of the amplified magnetic turbulence $P_w(x)$ rather than by
2271: the thermal pressure $\Pth(x)$. The situation is changed when the
2272: turbulence dissipation operates, because it dampens $P_w(x)$ and
2273: increases $\Pth(x)$, which reduces $\Msone$.
2274:
2275: Our getting such a high value of $\Msone$ is important because,
2276: just like \citet{KJG2002}, we have found in
2277: Section~\ref{injectionvsdissipation} that $\Msone \approx 10$ is a
2278: `breaking point' in the thermal leakage model of particle injection: the
2279: injection rate depends weakly on $\Msone$ when $\Msone \gtrsim 10$, but
2280: increases rapidly with decreasing $\Msone$ if $\Msone < 10$.
2281:
2282: %\clearpage
2283: \begin{figure}
2284: \epsscale{.50} \plotone{f6.eps}
2285: \caption{Illustration of momentum flux balance in the non-linear
2286: simulation with $\heatpar=0$ in the \coldismNoT\ case. The thin
2287: dashed line is the thermal pressure $\Pth(x)$, the thick dashed line -
2288: the CR pressure $\Pcr(x)$, the dotted line is the magnetic turbulence
2289: pressure $P_w(x)$, the dash-dotted line is the dynamic pressure
2290: $\rho(x) u^2(x)$, and the thick solid line is the sum of the four, the
2291: total momentum flux. All quantities are normalized to the far upstream
2292: value of the total momentum flux.
2293: \label{fig_momflx}}
2294: \end{figure}
2295: %\clearpage
2296:
2297: As an illustration for the above discussion, we
2298: show in Figure \ref{fig_momflx} the various constituents of the momentum
2299: flux across the shock with $\heatpar=0$ in the \coldism\ case.
2300: It's clear that upstream of the shock the dominant contributor to the
2301: momentum flux is the CR pressure $\Pcr(x)$ and, therefore, the precursor
2302: compression is mainly determined by $\Pcr(x)$. For this particular set
2303: of parameters, $\Pcr(x)$ results in a decrease in the flow speed $u(x)$
2304: by a factor of $\sim 5.4$ from its upstream value $u_0$ to the
2305: pre-subshock value $u_1$.
2306: The temperature in the precursor only increases adiabatically for
2307: $\heatpar=0$ and this results in an increase over the far upstream
2308: temperature $T_1/T_0 \approx 3.3$, thus reducing the local sonic Mach
2309: number to $\Msone \approx 44$.
2310: %The thermal pressure and the magnetic pressure remain low
2311: %for $x<0$ and don't come into play in determining the value of $\Msone$.
2312:
2313: The subshock compression ratio $\rsub$ is determined by the change in
2314: the different constituents of the momentum flux across the subshock.
2315: $\Pcr(x)$, although large, changes little across the
2316: subshock\footnote{
2317: %
2318: $\Pcr(x)$ is discontinuous at $x=0$ in our simulation
2319: due to the contribution to it of the particles that have crossed the
2320: shock only a few times. The jump in $\Pcr(x)$ in this case is small
2321: compared to the jump in thermal and magnetic pressures across the
2322: subshock, but can be significant for $\heatpar>0$, as seen in
2323: Figure~\ref{fig_partpress}.},
2324: %
2325: and what determines $\rsub$ is the change in $\Pth(x)$ and $P_w(x)$. The
2326: latter, as the plots in Figure \ref{fig_momflx} show, contributes
2327: significantly to the momentum flux. This is an important point because
2328: the values of $\Befftwo$ and $P_w(x)$ depend on $\rsub$ and $\rtot$ in a
2329: non-linear way making the traditional Rankine-Hugoniot relations
2330: inapplicable for determining $\rsub$.
2331: %
2332: Relation (\ref{rtotqesc}) can be used to
2333: iteratively calculate the resulting compression ratio $\rtot$, and it
2334: results in $\rtot=16$ and $\rsub=2.95$, as Table \ref{sumnl_cold}
2335: shows. As we can see, the effect of magnetic turbulence pressure tends to decrease
2336: $\rsub$ compared to the case $P_w(x) \ll \Pth(x)$ (as in the latter case
2337: one would expect $\rsub\approx 4$ for $\Msone \approx 44$).
2338:
2339: \clearpage
2340:
2341: \bibliographystyle{apj}
2342: %%\bibliography{c:/a_a_TOP/bibTeX/bib_DCE}
2343: \bibliography{bib_DCE}
2344:
2345:
2346: \end{document}
2347: