0808.1757/ms.tex
1: %\documentstyle[11pt,epsfig]{article}
2: \documentstyle[11pt, preprint]{aastex}
3: \textwidth=6.5in
4: \textheight=8.25in
5: 
6: \begin{document}
7: 
8: %\voffset=-0.3 true cm
9: %\hoffset=-1.8 true cm
10: 
11: \def\beq{\begin{equation}}
12: \def\eeq{\end{equation}}
13: \def\a{\alpha}
14: \def\t{\theta}
15: \def\r{{\it RHESSI}\,}
16: \def\ie{{\it i.e.}\,}
17: \def\eg{{\it e.g.}\,}
18: \def\ea{et al.\,}
19: \def\3he{$^3$He\,}
20: \def\he4{$^4$He\,}
21: \def\eq{equation\,\,}
22: \def\eqs{equations\,\,}
23: \def\cc{{\rm cm}^{-3}}
24: \def\g{\gamma}
25: \def\loss{{\rm loss}}
26: \def\Coul{{\rm Coul}}
27: \def\sy{{\rm synch}}
28: \def\scat{{\rm scat}}
29: \def\esc{{\rm esc}}
30: \def\eff{{\rm eff}}
31: \def\tr{{\rm cross}}
32: \def\ic{{\rm IC}}
33: \def\br{{\rm brem}}
34: \def\ac{{\rm ac}}
35: \def\min{{\rm min}}
36: \def\max{{\rm max}}
37: \def\tot{{\rm total}}
38: \def\rt{{\rm rad tot}}
39: \def\ln{{\rm ln}}
40: \def\Clog{\ln\Lambda}
41: \def\ltsim{\hbox{\rlap{\raise.5ex\hbox{$<$}}{\lower.7ex\hbox{$\sim$}}}}
42: \def\gtsim{\hbox{\rlap{\raise.5ex\hbox{$>$}}{\lower.7ex\hbox{$\sim$}}}}
43: 
44: \title{Particle Acceleration in Solar Flares and Enrichment of $^3$He and Heavy
45: Ions}
46: 
47: 
48: \author{Vah\'{e} Petrosian\altaffilmark{1,2,3}}
49: 
50: \altaffiltext{1}{Department of Physics, Stanford University, Stanford, CA, 94305
51: email; vahep@stanford.edu}
52: \altaffiltext{2}{Kavli Institute for Particle Astrophysics and Cosmology,
53: Stanford University, Stanford, CA 94305}
54: \altaffiltext{3}{Also Department of Applied Physics}
55: 
56: \begin{abstract}
57: We discuss possible mechanisms of acceleration of particles in solar flares and
58: show that turbulence plays  an 
59: important role in all the mechanism. It is also argued that  stochastic particle
60: acceleration by turbulent plasma waves is the most likely mechanism for
61: production of the high energy electrons and ions responsible for observed
62: radiative signatures of solar flares and for solar energetic particle or SEPs,
63: and that the predictions of this model agrees well with many past and recent
64: high spectral and temporal observations of solar flares. It is shown that, in
65: addition, the model explains many features of   SEPs  that accompany flares. In
66: particular we show that it can  successfully explain the observed extreme
67: enhancement, relative to photospheric values,  of $^3$He ions and the relative
68: spectra of $^3$He and $^4$He. It has also the potential of explaining the
69: relative abundances of most ions including the increasing  enhancements of heavy
70: ions with ion mass or mass-to-charge ratio.
71: 
72: \end{abstract}
73: 
74: \section{Introduction}
75: 
76: The aim of this paper is to investigate to what extent current theoretical
77: models of
78: acceleration of particles can explain the observed characteristics of solar
79: energetic particles (SEPs), in particular the extreme enhancement,
80: relative to photospheric values, of $^3$He ions in many flares, in particular in the so-called
81: impulsive (or He-rich) events. In the next section we  present a general
82: description
83: of the possible acceleration mechanisms for production of particles in solar
84: flare.  We show that plasma waves or turbulence play a major
85: role in all these processes. In \S 3 we describe the model of
86: stochastic
87: acceleration by plasma turbulence in some detail and review its successes in
88: describing observations of solar flares, in particular the radiative signatures
89: produced by accelerated electrons and protons. In \S 4 we address the
90: problem of observed enhancements
91: of \3he and heavy elements.  Here we first
92: present a brief review of some of the earlier models proposed for production of
93: these enhancements and then compare predictions of the stochastic acceleration
94: model with the observations. A brief summary and discussion is presented in \S
95: 5.
96: 
97: 
98: 
99: \section{Particle Acceleration}
100: \label{acceleration}
101: 
102: In this section we first compare various acceleration processes and stress the
103: importance of plasma wave turbulence  ({\bf PWT}) as an agent of
104: acceleration, and then describe the basic scenario and equations for treatment
105: of
106: these processes in the stochastic acceleration ({\bf SA}) model.
107: As described below there is growing evidence that PWT plays an
108: important role in acceleration of particles in general, and in solar flares in
109: particular. The two most commonly used acceleration mechanisms are the
110: following.
111: 
112: {\bf 1. Electric Field Acceleration:}  Electric fields parallel to
113: magnetic fields can accelerate charged particles.
114: For solar flare conditions the Dreicer field ${\bf E}_D = k_BT/(ev\tau_{\rm
115: Coul})\sim 10^{-5}$ V/cm. (Here $k_B, T, v$ and $e$ are the Boltzmann constant,
116: plasma temperature, particle velocity and charge,
117: and $\tau_{\rm Coul}$ is the mean collision time.)
118: Sub-Dreicer fields (${\bf E}\leq{\bf E}_D$) can only accelerate particles up to
119: 10's of keV (for a typical flare
120: length scale
121: $L\sim10^9$~cm)  which is far below the  10's of MeV electrons or $>10$ GeV
122: protons
123: required
124: by observations%
125: \footnote{Dreicer field will be large if the
126: resistivity and density are
127: anomalously high (Tsuneta 1985; Holman 1996a, b).}.  Super-Dreicer fields,
128: which seem to be present in many simulations of reconnection (Drake, 2006, see
129: also
130: Cassak \ea 2006; Hoshino 2006, see also Zenitani \& Hoshino, 2005),
131: accelerate particles at a rate that is faster than $\tau_{\rm Coul}$. This can
132: lead to a runaway and an unstable electron distribution which, as shown
133: theoretically,
134: by laboratory experiments and by the above mentioned simulations,
135: most probably will give rise to PWT (Boris et al.
136: 1970, Holman 1985)%
137: \footnote{Note also that  production of a broad
138: power-law requires a wide range of  potential drops
139: (Litvinenko, 2003).}. {\it In summary the electric fields arising as a result of
140: reconnection cannot be the sole agent of acceleration, but may produce an
141: unstable particle momentum distribution which will produce PWT that can then
142: accelerate particles.}
143: 
144: {\bf 2. Fermi Acceleration:} Nowadays this process has been divided into two
145: kinds.
146: In the original Fermi process particles of velocity $v$ scattering
147: randomly with a (pitch angle cosine $\mu$) diffusion rate $D_{\mu\mu}$ by moving
148:  scattering agents with a velocity $u$
149: gain energy at a rate proportional to $(u/v)^2D_{\mu\mu}$.
150: This, known as a {\it second order Fermi process}, is what we call {\it
151: stochastic acceleration}
152: ({\bf SA}). {\it For solar flares the most likely agent for scattering is PWT.}
153: An alternative process is what is commonly referred to as {\bf Shock}
154: acceleration.
155: Because particles
156: crossing the shock (or any kind of flow convergence) gain energy at a rate which
157: is
158: $\propto u_{sh}/v$, this is known as the {\it first order Fermi} process.
159: Ever since the demonstration by several authors that a very simple version of
160: this process leads to a power law spectrum that agrees approximately with
161: observations of the cosmic rays, shock acceleration is commonly invoked in space
162: and astrophysical
163: plasmas. However, this simple model, though very elegant, has many shortcomings
164: Specially
165: when applied to non-thermal radiation processes.
166: 
167: {\bf Shock} acceleration, as commonly used, requires injection of (sometimes
168: high energy)
169: particles and is unable to accelerate low energy background particles.
170: The simple results also break down when  one considers
171: more realistic models. For example, inclusion of losses
172: (Coulomb at low energies and synchrotron at high) or influence of
173: accelerated particles on the shock structure (see \eg Ellison \ea, 2005: Amato
174: \& Blasi 2005)
175: cause breaks and generally deviations from a simple power law.
176: More importantly, a {\it shock by itself cannot accelerate particles}
177: and requires some
178: scattering agents (most likely PWT) to cause repeated passages of the particles
179: through the shock front
180: (thus the name Fermi).  The rate of
181: energy gain is governed by the scattering rate $D_{\mu\mu}$.
182: This is the case for the so-called quasi-parallel shock, but  shocks
183: are most likely quasi-perpendicular in which case, as pointed out by Jokipii
184: (\eg 1987),
185: the particles may drift along the surface of the shock and get accelerated more
186: efficiently. However, this model also suffers from the above mentioned
187: fundamental limitation and
188: requires injection of particles with velocities $v>u_{sh}\eta$,
189: where $\eta \gg 1$ is the ratio of parallel to perpendicular (to magnetic field)
190: diffusion coefficients. In a more recent simulation of this process Giacolone
191: (2005)
192: shows that a non-thermal tail can be produced from a thermal distribution.
193: However, this model includes
194: significant field fluctuations, which as advocated below can accelerate low
195: energy
196: particles more efficiently
197: than shocks. Thus, it is not obvious what are the
198: relative roles of shock vs SA here.
199: Moreover, for solar flares there is no
200: direct evidence for shocks near the LT
201: during the impulsive phase, and some of the features that make
202: acceleration of cosmic rays by shocks attractive are not present%
203: {\footnote{For example, acceleration by sub Alfv\'enic shocks is
204: questionable (see Kulsrud 2005), and the streaming instability,
205: which is a part of the traditional shock acceleration
206: can be partially suppressed in the presence of PWT (Yan \& Lazarian 2002;
207: Farmer \& Goldreich 2004).}.
208: 
209: {\bf Stochastic Acceleration} is favored because the PWT needed for scattering
210: can also accelerate particles stochastically with the rate $D_{EE}/E^2$,
211: so that shocks may not be always necessary. First, contrary to common belief,
212: Hamilton \& Petrosian 1992 and 
213: Miller \& Reames 1996 have shown that for flare conditions PWT can
214: accelerate the background particles to high energies within the desired
215: time.  More importantly, at low
216: energies or in strongly magnetized plasmas the acceleration rate $D_{EE}/E^2$
217: may exceed the scattering rate $D_{\mu\mu}$ (see Pryadko \& Petrosian, {\bf
218: PP97}).  {\it Thus,
219: for flare conditions low energy particles are accelerated more efficiently by
220: PWT than
221: by shocks}. We note, however, that in gradual flares there may be a
222: need for a second stage acceleration of SEPs by high coronal shocks%
223: \footnote{In practice, \ie mathematically, there
224: is little difference between the two mechanisms (Jones 1994), and the
225: acceleration by turbulence and  shocks can be combined (see below).}.
226: 
227: {\it Irrespective of which process dominates the particle acceleration, it is
228: clear
229: that PWT has a
230: role in all of them. Thus, understanding of the production of PWT and
231: its interaction with particles  is extremely important.}
232: Moreover, {\bf turbulence} is expected to be present in most astrophysical
233: plasmas
234: including solar flares and in and around interplanetary shocks, because the
235: ordinary and magnetic Reynolds numbers for these situations are
236: very large and may be the most efficient channel of energy dissipations in
237: non-equilibrium systems such as solar flares.
238: In recent years there has been a substantial progress
239: in the understanding of
240: MHD turbulence (Goldreich \& Sridhar, 1995, 1997; Lithwick \& Goldreich 2001;
241: Cho \& Lazarian 2002 and 2006). These advances provide new tools for a more
242: quantitative
243: investigation of turbulence and the role it plays
244: in solar flares.
245: 
246: %\input{acceleration.tex}
247: 
248: %\input{SAmodel.tex}
249: 
250: \section{Stochastic Acceleration and Solar Flares}
251: \label{SAandFlares}
252: 
253: 
254: \subsection{Basic Scenario}
255: \label{scenario}
256: 
257: The complete picture of a solar flare involves many phases or steps.
258: After a complex pre-flare buildup, the first
259: phase is the reconnection or the energy release process.  The final consequences
260: of
261: this released energy are the observed radiations (from long wavelength radio to
262: GeV gamma-rays), SEPs and CMEs.  Many processes are involved in
263: conversion of the released energy into radiation. As stressed above
264: we believe that PWT plays an important role in these processes.  We envision the
265: following
266: scenario.
267: Magnetic energy is converted into turbulence by the reconnection process above
268: corona
269: loops which we refer to as the acceleration site or the loop top ({\bf LT})
270: source (see Figure \ref{model}). The turbulence or waves generated on some
271: macroscopic scale (some fraction of reconnection site scale) undergo
272: two kind of interactions. The  first is {\it nonlinear wave-wave interaction}
273: causing
274: them to undergo dissipationless cascade to smaller scales. The second  is
275: {\it wave-particle interaction} which becomes more  important at smaller scales.
276: This
277: damps the turbulence. The lost energy goes into heating the background plasma
278: and/or accelerating particles into a non-thermal
279: tail.
280: The accelerated particles on the open field lines escape the Sun and are
281: observed as SEPs having undergone varied degree of scatterings and acceleration
282: (\eg by interplanetary shocks) during their transport to the Earth. The escaping
283: electrons may also produce type III radio bursts. The particles on the field
284: lines associated with the closed loops spiral down towards the photosphere
285: and produce the observed
286: non-thermal radiations
287: via their interactions with the
288: background particles (and fields) along the loop and primarily at the footpoints
289: ({\bf FPs}).
290: However, most of the energy of the non-thermal particles is lost  by collisions
291: causing heating and evaporation of the colder chromospheric plasma, which is
292: responsible
293: for most of the softer thermal radiation. This process,
294: described by the hydrodynamic equations, has a
295: time-scale comparable to the sound travel time, and
296: is somewhat decoupled from the acceleration-transport process that has
297: a much shorter time-scale. However, the evaporation can modulate the high
298: energy processes by changing the density and temperature in the
299: acceleration site.
300: 
301: \subsection{Formalism}
302: \label{equations}
303: 
304: The mathematical treatment of the processes involved in such a scenario is long
305: and complex. Below we give an outline of the important aspects.
306: 
307: \subsubsection{Kinetic Equations}
308: \label{kinetic}
309: 
310: {\bf The spectrum of PWT} $W({\bf k},t)$ is determined by the wave-wave and
311: wave-particle interactions. The cascade is evaluated  from the rates of
312: wave-wave interactions. For
313: example the three wave interactions can be presented as (see \eg Chandran 2005;
314: Luo \& Melrose 2006)
315: \beq\label{wave-wave}
316: \omega({\bf k}_1) + \omega({\bf k}_2) = \omega({\bf k}_3) \,\,\,\, {\rm and}
317: \,\,\,\, {\bf
318: k}_1+{\bf k}_2={\bf k}_3,
319: \eeq
320: where $\omega$ and ${\bf k}$ are the wave frequency and wavevector. The
321: interaction
322: rates can be represented by the wave diffusion coefficient $D_{ij}$. In
323: general, at large scales the  cascade
324: time $\tau_{\rm cas}\sim k^2/D_{ij}$ is shorter than the damping time
325: $\Gamma(k)^{-1}$ resulting from wave-particle interactions. Thus, the waves
326: undergo dissipationless cascade from the injection scale $k_\min\sim L^{-1}$
327: (with a rate ${\dot Q}^W({\bf k})$) till $k_\max$ where $\tau_{\rm
328: cas}\Gamma=1$, known as the {\it inertial range}. Beyond $k_\max$ the
329: spectrum of waves drops off more rapidly.
330: Adopting the
331: diffusion approximation (see \eg Zhou \& Matthaeus 1990), one
332: can obtain
333: the evolution of the spatially integrated wave spectrum $W({\bf k}, t)$
334: \beq
335: \\{\partial {W} \over \partial t}
336: = {\partial \over\partial k_i}\left[D_{ij}{\partial\over \partial
337: k_j}{W}\right] -
338: \Gamma({\mathbf k}){W} - {{W}\over T^{W}_{\rm esc}({\mathbf k})} + \dot{Q}^{W}.
339: \label{waves}
340: \eeq
341: We will assume that all the wave energy is absorbed so that there is no wave
342: escape, \ie $T^{W}_{\rm esc}\rightarrow\infty$.
343: 
344: The general equation for treatment of the {\bf particle acceleration and
345: transport} is the
346: Fokker-Planck equation
347: for the gyro-phase averaged particle distribution $f(t,s,E,\mu)$ as a function
348: of
349: distance $s$ along the magnetic field lines. This equation is simplified
350: considerably if one can adopt the
351: isotropic
352: approximation (or deal with the pitch-angle averaged distribution) and 
353: impose
354: the homogeneity condition (or deal with distribution integrated over the
355: acceleration region). These are reasonable assumptions for particles at the
356: flare site and  amount to determination of the evolution of the
357: energy spectrum
358: $N(E,t)=\int \int d\mu ds f(t,s,E,\mu)$:
359: \beq
360: {\partial N \over \partial t}
361:  =  {\partial \over \partial E}\left[D_{EE}{\partial N \over \partial E}
362:  - (A-\dot E_L) N\right]
363:  - {N \over T^p_{\rm esc}} +{\dot Q}^p,
364: \label{kineq}
365: \eeq
366: Here  $D_{EE}/E^2, A(E)/E$ describe  the energy
367: diffusion and direct acceleration rates. They are obtained from consideration of
368: the wave-particle interactions, which are often dominated by resonant
369: interactions, specially for low beta (magnetically dominated) plasma, such that
370: \beq\label{resonance}
371: \omega({\bf k})-k\cos\t v\mu = n\Omega/\gamma,
372: \eeq
373: for  waves propagating at an angle $\t$ with respect to the large scale magnetic
374: field,
375: and a particle of velocity $v$, Lorentz factor $\gamma$, pitch angle cosine
376: $\mu$ and gyrofrequency $\Omega$. Here the harmonic number  $n$ (not to be
377: confused with the density)
378: is equal to zero for transit
379: time damping ({\bf TTD}) process. For parallel propagating waves ({\bf PPWs})
380: $n=\pm 1$
381: but for obliquely propagating waves
382: $n= \pm 1, \pm 2, ...$\,.
383: ${\dot E}_L/E$ is the energy loss rate of the particles (due mainly to Coulomb
384: collisions and synchrotron losses) and $\dot{Q}^p$ and the terms with the
385: escape times $T_{\rm esc}$ describe the source and leakage of particles.  (For a
386: more detailed discussion leading to the above equations see Miller et al. 1996
387: and Petrosian \& Liu 2004, {\bf PL04}).
388: 
389: The above two kinetic equations are coupled by the fact that
390: the coefficients of one depend on the spectral distribution of the other.
391: Conservation of energy requires that the energy lost by the waves ${\dot
392: {\cal
393: W}_{\rm tot}}\equiv \int \Gamma({\bf k})W({\bf k})d^3k$ be equal to the energy
394: gained by the
395: particles from the waves;  ${\dot {\cal E}}=\int [A(E)-A_{\rm sh}]N(E)d E$.
396: Representing the
397: energy transfer rate between the waves and particles by
398: $\Sigma ({\bf k}, E)$ this equality implies that
399: \beq
400: \label{coefficients}
401: \Gamma({\bf k})= \int_0^\infty \d E N(E)\Sigma({\bf k}, E),\,\,\,\,
402: A(E)=\int_0^\infty d^3kW({\bf k})\Sigma ({\bf k}, E)+ A_{\rm sh}\ ,
403: \eeq
404: where we have added $A_{\rm sh}$ to represent
405: contributions of other (non-SA) processes affecting the direct acceleration, \eg
406: shocks.
407: 
408: The transport effects and further acceleration of the particles as they travel
409: in the interplanetary space can also be treated by similar equations.
410: 
411: \subsubsection{Dispersion Relations}
412: 
413: At the core of the evaluation of all the coefficients of the kinetic equations
414: described above lies the
415: plasma dispersion relation $\omega({\bf k})$ which determines the
416: characteristics of the waves that can be excited in the plasma, and the rates of
417: wave-wave and wave-particle interactions.
418: For example, in the MHD regime for a cold plasma
419: \beq
420: \label{mhddisp}
421: \omega=v_{\rm A}k\cos\t\,\,\,\,\,  {\rm and}\,\,\,\,\,  \omega=v_{\rm A}k
422: \eeq
423: for the Alfv\'en  and fast magneto-sonic waves, respectively, where $v_{\rm A}$
424: is the Alfv\'en velocity. Beyond the MHD regime a multiplicity of wave modes can
425: be
426: present and the dispersion relation is more complex and is obtained from the
427: following 
428: expressions:
429: \beq\label{dispgeneral}
430: \tan^2\theta = {-P(n_r^2-R)(n_r^2-L)\over (Sn_r^2-REL)(n_r^2-P)}
431: \eeq
432: where $n_r=kc/\omega$ is the refractive index, $S={1\over 2}(R+L)$, and
433: \beq\label{dispterms}
434: P= 1-\sum_i{\omega_{pi}^2\over \omega^2},\ \ \  R=1-\sum_i{\omega_{pi}^2\over
435: \omega^2}\left({\omega\over 
436: \omega+\epsilon_i\Omega_i}\right),\ \ \ {\rm and} \ \ \
437: L= 1-\sum_i{\omega_{pi}^2\over \omega^2}\left({\omega\over 
438: \omega-\epsilon_i\Omega_i}\right).
439: \eeq
440: Here $\omega_{pi}^2={4\pi n_i q_i^2\over m_i}$ and
441: $\Omega_i={|q_i|B\over m_ic}$ are the plasma and gyro frequencies,
442: $\epsilon_i={q_i\over |q_i|}$, and $n_i$, $q_i$, and $m_i$ are the  
443: density, charge, and mass of the background particles (electron, proton and
444: $\alpha$.)
445: Figure \ref{dispersion} shows the more complete picture of the dispersion
446: surfaces  (depicted by the
447: curves) obtained from the well known cold plasma expressions
448: along with the resonant planes
449: in  the ($\omega,\ k_\parallel,\ k_\perp$) space. Intersections
450: between the dispersion
451: surfaces and the resonant planes define the resonant wave-particle interactions.
452: 
453: \begin{figure}[hbtp]
454: \centerline{
455: \includegraphics[height=8.5cm,angle=270]{dispelect.ps}
456: \includegraphics[height=8.0cm,angle=270]{dispprot.ps}
457: }
458: \caption
459: {\scriptsize
460: Dispersion relation (curves) and resonance condition (flat) surfaces
461: for a cold fully ionized H and He (10\% by number) plasma with
462: $\beta_{\rm A}=v_{\rm A}/c=0.012$
463: showing the regions
464: around the electron (left) and proton (right) gyro-frequencies. Only
465: waves with positive $k_\|, k_\perp$ (or $0<\t<\pi/2$) are shown.
466: The mirror image with respect to the ($\omega,\ k_{\perp}$) plane
467: gives the waves propagating in the opposite direction. From high to
468: low frequencies, we have one of the electromagnetic branches (green),
469: upper-hybrid
470: branch (purple), lower-hybrid branch, which also includes the whistler waves
471: (pink),
472: fast-wave branches (yellow), and Alfv\'{e}n
473: branch (black). The effects of a finite temperature modify these curves at
474: frequencies $\omega\sim kv_{\rm th}$, where $v_{\rm th}=\sqrt{2k_{\rm B}T/m}$ is
475: the thermal velocity (see \eg Andr\'{e} 1985). The
476: resonance surfaces are for electrons with $v=0.3c$ and $|\mu|=1.0$ (left: upper
477: $n=1$, lower $n=0$) and
478: $^4$He (right: middle $n=1$) and $^3$He (right: upper $n=1$) ions with
479: $|\mu|=1.0$ and $v=0.01c$. The
480: resonance surfaces for the latter two are the same when $n=0$ (right: lower).}
481: \label{dispersion}
482: \end{figure}
483: 
484: The above dispersion relations are good approximations for low beta plasmas,
485: \beq
486: \label{beta}
487: \beta_p=2(v_s/v_{\rm A})^2=8\pi nk_{\rm B}T/B^2=3.4\times 10^{-2}(n/10^{10}
488: {\rm cm}^{-3})(100 {\rm G}/B)^2(T/10^7{\rm K})\ll 1,
489: \eeq
490: where $v_s=\sqrt{k_{\rm B}T/m_p}$ is the sound speed. For higher beta
491: plasmas,
492: \eg  at higher temperatures, these relations
493: are modified, specially for higher frequencies $\omega\sim kv_{\rm th}$, where
494: $v_{\rm th}=\sqrt{2k_{\rm B}T/m}$. For example, in the MHD regime, in
495: addition
496: to the Alfv\'en mode one gets fast and slow modes with the  dispersion relation
497: (see \eg Sturrock 1994)
498: \beq
499: \label{mhddisp2}
500: (\omega/k)^2={1\over 2}\left[(v_{\rm A}^2+v_s^2)\pm \sqrt{v_{\rm
501: A}^4+v_s^4-2v_{\rm A}^2v_s^2\cos {2\t}}\right],
502: \eeq
503: and the more general dispersion relation is modified in a more complicated way
504: (see \eg Andr\'{e} 1985 or Swanson 1989). The finite temperature imparts an
505: imaginary part to the wave frequency that gives the (Landau) damping rate of the
506: waves (see \eg Swanson 1989 or Pryadko \& Petrosian 1998, 1999 and Cranmer \&
507: van Ballegooijen 2003 for application
508: to solar flare conditions). In general, these rates and the
509: modification of the dispersion relation are known for Maxwellian (sometimes
510: anisotropic) energy distribution of the plasma particles. For non-thermal
511: distributions the damping rates can be evaluated as described in Petrosian, Yan
512: \& Lazarian (2006) using the
513: coupling described in equation  \ref{coefficients}. If the damping due to
514: non-thermal particles is important then the wave and particle kinetic equations
515: (\ref{waves}) and (\ref{kineq}) are coupled. This will be important in large
516: intense flares where a substantial fraction of particles are accelerated into a
517: non-thermal
518: tail. However, most often the damping rate is
519: dominated by the background thermal particles so that the wave and non-thermal
520: particle kinetic
521: equations decouple. This simplifies matters considerably. One can then use the
522: imaginary part of the wave frequency for evaluation of the damping rate using
523: the well-known hot plasma dispersion relation. Moreover, the 'thermal' effects
524: change the real part of the wave frequency only slightly so that
525: often the real part (and the particle diffusion coefficients) can be evaluated
526: using the simpler cold plasma dispersion relation depicted in Figure
527: \ref{dispersion}.
528: 
529: 
530: 
531: \subsubsection{Some Model Results} 
532:  
533: We now demonstrate some of the basic characteristics
534: of our scenario which can be compared with observations. The complete treatment
535: of the problem requires solution of the coupled wave particle kinetic equations
536: for particles at all energies and pitch angles and waves propagating at all
537: angles. In principle, given the plasma density, magnetic field, temperature
538: ($n,\
539: B,\ T$), the geometry of the
540: region
541: (represented by a size $L$ here), the rate and scale of injection  of
542: turbulence $\dot{Q}^W({\bf k}, t)$, one can evaluate the coefficients of
543: equations
544: (\ref{waves}) and (\ref{kineq}) and solve the coupled kinetic equations for
545: determination
546: of the resultant distributions $N(E,t)$ and $W({\bf k},t)$. However,
547: under certain circumstances some simplifications are possible.
548: As mentioned above in most cases the damping by non-thermal particles can be
549: ignored and the equations are decoupled. The wave spectrum then can be
550: approximated by a power law in the inertial range ($W\propto k^{-q}, k_{\rm
551: min}<k<k_{\rm max}$) with a steeper falling spectrum beyond this range. The
552: exponent $q$ will be in the range expected in a Kolmogorov or Kraichnan cascade.
553: Furthermore, as shown in Petrosian \ea (2006), the upper limit $k_{\rm max}$
554: depends strongly on the angle of propagation $\t$ with only quasi-parallel
555: propagating waves {\bf PPWs} reaching small enough scale to be in resonant with
556: low energy
557: particles. Thus, the initial acceleration of thermal particles will be dominated
558: by such waves. Consequently, in what follows here and the next section we assume
559: PPWs with the above spectrum. The following
560: results are based on PL04, where  more details can be found. 
561: 
562: \begin{figure}[htbp]
563: \leavevmode
564: \centerline{
565: %\vspace{-1.0cm}
566: \includegraphics[width=7.5cm]{ppwdispep.eps}
567: \includegraphics[width=7.5cm]{times.eps}
568: }
569: \caption{\scriptsize
570: {\bf Left panel:} The dispersion relation for PPWs (curves)  and the resonant
571: relation (lines, from eq.[\ref{resonance}]) for electrons (top) and protons
572: (bottom, expanded near the origin). Note that the resonant lines start (at
573: $k=0$) from the gyro frequency of the particle (divided by the Lorentz factor
574: $\gamma\sim 1$).
575: {\bf Right panel:} Time scales for a SA acceleration model for solar flare
576: conditions and an
577: assumed
578: spectrum of PWT. Model parameters are indicated in the figure
579: ($\tau_p^{-1}\propto \Omega_e{\cal W}/B^2$ is a
580: characteristic wave-particle interaction rate and $\alpha\propto \sqrt{n}/B$).
581: The acceleration, escape, and
582: loss times are indicated by the dotted, dashed, and solid curves, respectively.
583: The thick
584: (red) curves are for electrons and thin (blue) curves are for protons.}
585: \label{timescales}
586: \vspace{-0.5cm}
587: \end{figure}
588: 
589: 
590: 
591: The left panel of Figure \ref{timescales} shows the
592: dispersion relation for PPW and the resonant conditions (straight lines) for an
593: electron and a proton of the specified energy and pitch angle cosine. The main
594: parameter determining
595: these curves and the resulting resonances (intersections of the lines with
596: the curves) is the ratio of the plasma to gyro frequencies
597: \beq\label{alpha}
598: \alpha=\omega_{pe}/\Omega_e=(m_e/m_p)^{1/2}/\beta_{\rm A}=3.2(n/10^{10}
599: \cc)^{1/2}(100 {\rm G}/B).
600: \eeq
601: The right panel of Figure \ref{timescales} shows the timescales associated with
602: the terms in equation (\ref{kineq}), {\it i.e.} loss
603: time
604: $\tau_{\rm loss}=E/{\dot {E}_L}$, diffusion time $\tau_{\rm diff} \sim
605: E^2/D_{EE}$, acceleration time $\tau_{\rm a}=E/A(E)$, and escape time% 
606: \footnote{This is an approximate
607: relation valid for scattering time $\tau_{\rm scat}\ltsim D_{\mu\mu}^{-1}>$
608: longer and
609: shorter than $T_{\rm cross}\sim L/v$, the time for particles to cross the
610: turbulent region
611: of size $L$ (see PL04).}$T_{\rm esc}= T_{\rm cross}(1+T_{\rm cross}/\tau_{\rm
612: scat})$.
613: 
614: \begin{figure}[htbp]
615: \leavevmode
616: \centerline{
617: \includegraphics[width=7.5cm]{epspectra.eps}
618: \includegraphics[width=7.5cm]{espectra_taup.ps}
619: }
620: \caption{\scriptsize
621: {\bf Left panel:}
622: The spectra of accelerated electrons and protons corresponding to the timescales
623: shown in Figure \ref{timescales} (right). The dotted
624: curves give the spectra at the acceleration site (LT) and the dashed curves give
625: the thick target equivalent
626: spectra of escaping particles. The solid (black) line
627: gives the shape of the background (injected) particle distribution. The hatched
628: areas show the energy ranges for production of  frequently observed ranges of
629: HXRs (for electrons) and gamma-ray lines (for protons). Spectral features occur
630: at
631: the intersections and divergences of these timescales (circle and square signs
632: shown here and in Fig. \ref{timescales}; right).
633: {\bf Right panel:} The dependence of electron spectrum on the  ${\cal
634: W}\propto\tau_p^{-1}$.
635: Higher ${\cal W}$'s produce harder spectra
636: and more acceleration vs heating (see PL04 ).
637: }
638: \label{spectra}
639: \vspace{-0.5cm}
640: \end{figure}
641: 
642: {\it The spectral characteristics} of the particles
643: arise from the interplay among the
644: timescales shown above.
645: Figure \ref{spectra} (left) shows
646: the corresponding spectra  of electrons and protons at the LT
647: ($N_{\rm LT}$, dotted) and the equivalent FP thick target spectra (dashed lines;
648: see \eg Petrosian 1973, Park, Petrosian \& Schwartz  1997)
649: \beq\label{thick}
650: N_{\rm FP}(E)={\dot E}^{-1}_L\int_0^E N_{\rm LT}(E')T_{\rm esc}^{-1}(E')dE'.
651: \eeq
652: Spectral breaks
653: occur at energies (circles and squares) when different terms
654: become important. As evident, in general, SA by PWT  produces a
655: ``quasi-thermal''
656: and a harder non-thermal component; \ie it both ``heats" the plasma and
657: accelerates
658: particles. As also seen in the right panel of Figure \ref{spectra}, the relative
659: strength of the two components depends
660: sensitively on the intensity ${\cal W}\propto \tau_p^{-1}$ of the PWT. For
661: higher densities of turbulence the spectra are harder. In addition,
662: the LT spectra are dominated by the
663: thermal component with a steep non-thermal component, while the FP  spectra
664: are harder with little or non-thermal part.  These agree with the observed
665: features discussed below.
666: For low ${\cal W}$, \eg during the build-up
667: and decay phases of a flare, one expects only plasma heating with essentially no
668: non-thermal component (see below and PL04 ).
669: %differences between the LT and FP sources,
670: %and can  explain the observed pre- and post-impulsive phase behaviors and the
671: %impulsive phase soft-hard-soft evolution
672: 
673: 
674: \subsection{Application to Solar Flares}
675: \label{observe}
676: 
677: Ultimately observations must determine the validity of the models.
678: As described below there have been  extensive comparisons of the SA model with
679: observations. To our knowledge, there have not been similarly detailed
680: comparisons for electric field or shock acceleration.
681: 
682: 
683: \subsubsection{Radiative Signatures of Electrons}
684: \label{electron}
685: 
686: {\bf Spatial Structure and Evolution:}
687: The most direct evidence for the presence of
688: PWT is observations by {\it Yohkoh} of distinct LT impulsive HXR emission
689: (Masuda et al. 1994, Masuda 1994), which  have been
690: shown to be common to almost all {\it
691: Yohkoh} (Petrosian, et al. 2002) and {\it RHESSI} (Liu et al. 2004) flares.
692: Petrosian \&
693: Donaghy (1999) have also shown that these observations require an enhanced pitch
694: angle scattering to confine the particles near the LT. Coulomb collisions cannot
695: be
696: this agent because of high losses they entail.
697: The most likely scattering agent is PWT, which can also accelerate
698: particles. Figure \ref{model} (left) shows a  cartoon of the
699: reconnection and acceleration site
700: for a flaring loop with the red foam representing the PWT produced
701: during reconnection. As reconnection proceeds larger closed loops
702: are formed below this site. This simple picture then predicts  a gradual rise
703: of the LT
704: source accompanied with a continuous increase in separation of the FPs. Most
705: strong flares are
706: complex and involve multiple (or arcade) of loops (\eg July 23, 2002 flare;
707: Krucker \ea 2003), which obscures the simple single loop behavior.
708: In  weaker flares, on the other hand, this behavior is lost in the noise.
709: {\it RHESSI} has detected a strong X-class flare (Nov. 3, 2003)
710: consisting of a single
711: loop which shows exactly this behavior (Liu \ea 2004, see Fig. \ref{model},
712: middle). Similar motion but only for the LT source has been seen in
713: several other {\it RHESSI} flares (see \eg Sui \& Holman 2003).
714: In rare cases, specially when the FP sources are weak or occulted by the sun,
715: one can see a double LT source (Fig. \ref{model} right, from Liu 2007,
716: see also Sui \ea 2005) as envisioned by the cartoon on the left.
717: 
718: 
719: \begin{figure}[hbtp]
720: \leavevmode
721: \centerline{
722: \hspace{0.7cm}
723: \includegraphics[width=5.5cm]{f3a.eps}
724: \hspace{-0.5cm}
725: \includegraphics[height=6.0cm, origin=c, angle=90]{f3b.eps}
726: \hspace{-1.0cm}
727: \includegraphics[width=6.7cm,height=6.0cm]{f3c.eps}
728: %\includegraphics[width=7.0cm,height=6.5cm,angle=90]{f3b.eps}
729: %\includegraphics[width=5.5cm,height=5.0cm]{f3c.eps}
730: }
731: \caption{\scriptsize
732: {\bf Left panel:} A schematic representation of the reconnecting field forming
733: closed
734: loops and open coronal field lines. The red foam represents PWT. {\bf Middle
735: panel:}
736: Temporal evolution of LT and FP HXR sources of Nov. 3, 2003 flare.
737: The symbols indicate the source centroids and the colors show the time with a 20
738: sec
739: interval, starting from black (09:46:20 UT) and ending at red (10:01:00 UT) with
740: contours for the last time. The curves connect schematically
741: the FPs and the LT sources for three different times showing
742: the expected evolution for the model at the left; from Liu \ea (2004). {\bf
743: Right panel:}
744: Image of flare of April 30, 2002, with occulted FPs
745: showing an elongated LT source with two distinct peaks as expected from
746: the model in the left (see Liu et al. 2008). The curves representing the
747: magnetic lines (added by
748: hand)
749: show the presumed occulted FPs below the limb (red line); from Liu (2007).
750: }
751: \label{model}
752: %\vspace {-0.5cm}
753: \end{figure}
754: 
755: 
756: 
757: {\bf Spectral Characteristics:}
758: Observations over a broad energy range (\eg Marschh\"{a}user et al.  1994,
759: Dingus et al. 1994, and
760: Park, Petrosian, \& Schwartz 1997) show several spectral breaks. Petrosian et
761: al. (1994) have shown that these features
762: present in the so-called
763: electron-dominated flares
764: cannot be due to transport or optical depth effects
765:  and are natural
766: consequences of the SA by PWT (Park et al. 1997).
767: Early high resolution observations (Lin et al.  1981)
768: showed spectral steepening below a few tens of keV, which is also consistent
769: with the SA model
770: (Hamilton \& Petrosian 1992). {\it RHESSI}'s imaging spectroscopy has clarified
771: this situation
772: considerably. In general, during the impulsive phase the spectra are dominated
773: by a non-thermal
774: component, with the LT always softer than the FPs, and there is a
775: quasi-thermal component (mostly from the LT)
776: occurring sometimes before the impulsive phase and always
777: during the decay phase, which is in concordance with the features of the
778: accelerated electrons described above.
779: Figure \ref{ionspectra} shows these observed characteristics of the LT and FP
780: sources 
781: for two typical {\it RHESSI} x-ray flares
782: fitted with theoretical spectra from the SA model.
783: 
784: \begin{figure}[hbtp]
785: \leavevmode
786: \centerline{
787: \includegraphics[width=7.5cm, height=5.5cm]{photonspec1.ps}
788: \includegraphics[width=7.5cm, height=5.5cm]{photonspec2.ps}
789: }
790: \caption{\scriptsize
791: {\bf Left panel:} A fit to the total (black), FP (green) and LT (red) spectra of
792: the 
793: first HXR pulse of the Sep.
794: 20th 2002 flare observed by {\it RHESSI}. Model parameters are indicated in the
795: figure. 
796: The dashed and dotted
797: lines are model spectra from equations (\ref{kineq}) and (\ref{thick}) showing
798: the 
799: presence of a
800: quasi-thermal (LT) and a non-thermal (mostly FPs) component. The solid line
801: gives the 
802: sum of the two. The blue
803: dashes indicate the level of the background radiation.
804: {\bf Right panel:} Same as the left panel but for the second HXR peak. Note
805: that 
806: compared with the model parameters
807: for the first HXR pulse, both the gas density and the turbulence energy density 
808: increase during the second pulse,
809: resulting more thermal emission and harder photon spectra.
810: } 
811: \label{ionspectra}
812: \end{figure}
813: 
814: {\bf Spectral Evolution:} It has also been known for
815: sometime that some observations disagree with the model where  \underline{all}
816: the released energy
817: goes directly to non-thermal electrons during {\it the impulsive phase}.
818: Among these are precursor soft X-ray emission referred to as {\it preheating}
819: and
820: a slower than expected temperature decline in the decay phase (see, {\it e.g.}
821: McTiernan et al.  1993)%
822: \footnote{The expected Neupert (1968) relation also does not seem to
823: hold (see Dennis \& Zaro 1993; Veronig et al. 2005).}.
824: In several {\it RHESSI} limb flares Jiang \ea (2006) find that during the decay
825: phase the
826: (resolved) LT
827: source continues to be confined; it does not extend to the FPs as one would
828: expect
829: if the evaporation
830: fills the whole loop with hot plasmas.
831: %Figure \ref{spectra} (right) shows the
832: %temporal evolution of the temperature and emission measure from spectral
833: %fitting,
834: %and the density from the
835: %observed source volume evolution  for the 09/20/2002 flare. The bottom  shows
836: %that the
837: Moreover,  the observed energy decay rate is calculated to be
838: higher than the radiative cooling rate
839: but much lower than the (Spitzer 1961) conduction
840: rate. The latter and the confinement of the LT source require suppression of the
841: conduction and a continuous energy input.
842: PWT can be the agent for both. It can reduce the scattering mean
843: free path and energize the plasma continuously during the decay phase (albeit at
844: a
845: diminishing rate).
846: 
847: The observed spectral  breaks and the temporal evolution also agree with the
848: model results discussed above. During the build-up and decay (\ie pre- and
849: post-impulsive)
850: phases when the level of PWT is low it only manages to ``heat" the plasma with
851: little
852: acceleration into a non-thermal tail. Particle acceleration begins in earnest
853: during
854: the impulsive phase when the rate  of reconnection and production of PWT rises
855: and then falls
856: after reaching high levels. During this rise and fall the spectra undergo
857: soft-hard-soft
858: evolution as observed in most flares.
859: 
860: \begin{figure}[hbtp]
861: \leavevmode
862: \centerline{
863: %\hspace{0.7cm}
864: \includegraphics[width=7.0cm]{epspectra1.eps}
865: %\hspace{-0.5cm}
866: \includegraphics[width=7.0cm]{epalpha.ps}
867: }
868: \caption{\scriptsize
869: {\bf Left panel:}
870: Same as left panel of Figure \ref{spectra} but for a different $\alpha$ showing
871: the dependence of relative electro and proton spectra on $\alpha$.
872: {\bf Right panel:} Variation with $\alpha$ of the ratio of the energies of the
873: accelerated electrons and protons in the spectral energy range shown by the
874: shade region in the left panel.
875: }
876: \label{epratio}
877: \vspace {-0.5cm}
878: \end{figure}
879: 
880: \subsubsection{Radiative Signatures of Protons and Ions}
881: 
882: Observations  of gamma-ray lines in larger solar flares provide
883: evidence for acceleration of protons and ions. The SA model has been
884: the working hypothesis
885: here as well  (see works by Ramaty and collaborators earlier, and Murphy and
886: Share recently). For these
887: flares
888: the ratio of energy content of accelerated electrons to
889: protons
890: (in the hatched range of Fig. \ref{spectra}, left)
891: ranges from 100 to 0.03
892: (Mandzhavidze \& Ramaty 1996);  this ratio is  larger in electron dominated
893: flares.
894: On the other hand, simple theoretical models, for example
895: acceleration by Alfv\'{e}n waves commonly used, favor a much smaller ratio.
896: PL04  show that with a more complete treatment of the
897: wave-particle interaction a barrier appears for proton acceleration,
898: which reduces the number of accelerated protons dramatically,
899: in a way that the above ratio is very sensitive to the plasma parameter
900: $\alpha$ in equation (\ref{alpha}).
901: Figure \ref{epratio} shows this dependence. The left panel is same as the left
902: panel in Figure \ref{spectra} but for a smaller $\alpha$. The right panel shows
903: the variation with $\alpha$ of the ratio of the energy of electron to protons
904: (in the respective shade region).
905: One consequence of this is
906: that the proton acceleration will be more efficient in larger loops, where the
907: $B$ field is
908: weaker, and at late phases, when  due to evaporation $n$ is higher.
909: This can explain the offset of the centroid of the gamma-ray line emissions from
910: that of HXRs seen by {\it RHESSI} (Hurford et al.  2003).
911: It can also account for the observed delay of nuclear line emissions relative
912: to HXR emissions seen in some flares (Chupp et al.  1990).
913: This trend is also true for other ions because the acceleration barrier
914: moves to lower energies with decreasing $\alpha$.
915: %This effect is shown on
916: %Figure \ref{He} (left) for $^4$He and will be true for all ions with the same
917: or
918: %similar charge-to-mass ratios.
919: In general, higher densities, lower magnetic fields favor
920: acceleration of ions
921: versus the electrons and can explain the diversity in their relative observed
922: values both at the flare site and in the interplanetary space.
923: 
924: \subsubsection{SEP Spectra and Abundances}
925: 
926: It is commonly believed that the observed relative abundances of ions in SEPs
927: favor a SA model
928: ({\it e.g.}
929: Mason et al. 1986 and Mazur et al. 1992). More recent observations have
930: confirmed
931: this picture (see
932: Mason et al. 2000, 2002, Reames et al. 1994 and 1997, and Miller 2003).
933: There are similarities and
934: differences among the spectra and abundances of different ions
935: from event to event.
936: One of the most vexing problem of SEPs has been the  enhancement of $^3$He
937: in the so-called {\it impulsive events}, which sometimes can be $3-4$ orders of
938: magnitude above the
939: photospheric value. As stated at the outset these events, which we prefer to
940: call
941: \3he-rich rather than impulsive, is the focus of thees proceedings.  There has 
942: been many attempts to  explain this
943: and the related enhancements of heavy ions observations. This will be the
944: subject of the next sections.
945: 
946: \section{Enhancement of \3he and Heavy Ions}
947: 
948: As stressed in the previous chapters one of the most puzzling aspects of
949: observations of SEPs is the enhancement of \3he and heavy ions  relative to
950: their photospheric
951: values in many of the so-called  impulsive events. Over the years there has been
952: many attempts to account for these observations. In this section we first
953: present a
954: brief
955: review of past attempts at tackling this problem and then describe how well the
956: SA
957: model described above can account for these aspects of solar flares.
958: 
959: 
960: \subsection{Past Works}
961: 
962: 
963: Most of the proposed models, except the  Ramaty and Kozlovsky (1974) model based
964: on  spalation (which has many problems), rely on resonant wave-particle
965: interactions
966: to produce the observed enhancement. The  earliest theoretical
967: models addressing this issue are those of Ibragimov and Kocharov (1977) and Fisk
968: (1978).
969: 
970: \begin{itemize}
971: 
972: \item
973: 
974: Ibragimov and Kocharov claim that under the influence of Langmuir turbulence the
975: plasma ion temperature will increase linearly in time with a rate proportional
976: to the momentum diffusion coefficient. Citing another paper they also claim that
977: the diffusion rate will be proportional to the Coulomb charge times the
978: gyrofrequency which favors heating of \3he relative to protons and \he4. A
979: second stage of acceleration with a velocity or energy threshold for all
980: particles will then accelerate a larger fraction  of \3he ions than \he4 and
981: proton. The parameters of the model can be adjusted to obtain the observed
982: enhancements.
983: \item
984: 
985: Fisk's model relies on the fact that  \3he  can interact more efficiently
986: than
987: \he4 with (electrostatic hydrogen-cyclotron) waves having frequency between
988: proton and \he4 gyrofrequencies. Damping of these waves preferentially heats
989: \3he  ions. There are two major shortcomings in this model. The main
990: difficulty
991: here is that to have sufficient overheating one requires excitation of waves
992: close to \3he  gyrofrequency (=4/3 of \he4) which can be produced if electron
993: to
994: proton temperature ratio and/or the \he4 to hydrogen ratio are large. Again,
995: this model does not answer the full question of production of enhanced
996: population of non-thermal \3he ions. It leaves this to a secondary process which
997: also
998: must satisfy the requirement of having a threshold velocity for acceleration
999: near
1000: the thermal velocity of the heated \3he ions.  Fisk also
1001: tries to explain the enhancement of the heavy elements the same way but relying
1002: on the resonant interactions with the second gyro-harmonics. This aspect also
1003: may not be correct.
1004: 
1005: \item
1006: 
1007: Temerin and Roth (1992) use essentially same idea but with
1008: electromagnetic hydrogen-cyclotron waves. They carry out numerical simulation
1009: based on somewhat artificial model (waves moving down the variable field lines
1010: and \3he  ions moving up that interact when the resonant condition is
1011: satisfied).
1012:  They also produce some heating of \3he  ions but do not discuss the spectrum
1013: nor
1014: the expected ratio of \3he/\he4. Again, like  in the Fisk model first  and
1015: second harmonics
1016: ($n=1$ and 2) are invoked for \3he  and heavy ion heating, respectively.
1017: However, there is no estimate of heavy ion fluxes. Curiously, Fisk is not
1018: acknowledged except in passing mentioning only the shortcoming  mentioned above.
1019: 
1020: \item
1021: 
1022: Miller and Vi\~nas (1993) expand on Temerin and Roth work by
1023: inclusion of more waves and more-realistic plasma conditions. Most of this
1024: paper is devoted to calculating the growth rate  and dispersion relation of
1025: waves excited by a thermal, isotropic ``beam" of electrons superimposed as  a
1026: bump in a somewhat cooler thermal plasma consisting of electrons,
1027: protons and alpha particles. The acceleration of particles is carried out using
1028: orbit calculations for test particles. As in the above works it merely leads to
1029: some heating of iron and \3he  ions. It is then stated that this heating is
1030: more
1031: than what they expect to occur for other ions (\eg CNO) and \he4, respectively.
1032: The required ``beam" densities are much higher than those required for
1033: describing
1034: the Type III radio emission. (These authors provide a good summary and analysis
1035: of earlier works).
1036: 
1037: \item
1038: 
1039: Zhang (1995) uses a more refined version of  the Fisk model and evaluates the
1040: degree of overheating of \3he ions relative to \he4. In addition he includes a
1041: simple calculation of subsequent acceleration by a ``Fermi" process and derives
1042: a general spectrum. However, these spectra are not compared with the observed
1043: ones; they are essentially exponential in form and will not fit well to the
1044: observations. Instead he evaluates the abundance ratios and plots these versus
1045: some important parameters which can be compared with the gross features of the
1046: observations.
1047: 
1048: 
1049: \item
1050: 
1051: Paesold, Kallenbach and Benz (2003)  rely on a somewhat different
1052: aspect of the acceleration process. In their model
1053: waves
1054: are excited by anisotropic distribution of electrons (so-called Fire Hose
1055: Instability) with a growth rate  that is slightly higher at the \3he
1056: gyrofrequency
1057: than that of \he4. It is then claimed that this can lead to a much larger
1058: (exponential)
1059: difference in the wave energy density at the corresponding frequencies. With a
1060: larger wave density at its disposal \3he  acceleration proceed more rapidly. The
1061: acceleration method is used is the same as in Miller and Vi\~nas discussed
1062: above. However, unlike the above works this paper addresses the spectral
1063: differences between \3he  and \he4, but there is no comparison with actual
1064: observed spectra. This model also predicts that there should be i) many more
1065: accelerated protons (which is swept under the rug invoking some energy
1066: limitation) and ii) a \underline{suppressions} of iron and other heavier ions
1067: (whose
1068:  $Q/A$  is less than that of \he4 or CNO) instead of the observed
1069: enhancements.
1070: 
1071: \end{itemize}
1072: 
1073: In summary the common theme among all these models is that they use the
1074: difference between the $Q/A$ of \3he and \he4 and require some non-Maxwellian
1075: electron distribution to excite the waves that then preferentially heat
1076: \3he ions. Some also require very special conditions.
1077: The outcomes of all these works are in general some differential heating with
1078: the actual acceleration
1079: often relegated to a subsequent process. There is little or no comparison with
1080: observed spectra and every model requires some additional factor to explain the
1081: enhancements observed for iron and heavier ions.
1082: 
1083: In contrast to this the model proposed by Liu, Petrosian and Mason (2004, 2006)
1084: based on the SA by PWT (PL04) treats heating and acceleration as a single
1085: process, can
1086: reproduce some of the observed spectral features, and can account for the range
1087: of the observed
1088: enhancement. On the other hand, at this stage of its development this model does
1089: not address the details of the production of the waves or turbulence and relies
1090: on injection, presumably during the reconnection process,  and the subsequent
1091: cascade and damping of the waves. Some of the salient results of this model are
1092: described next.
1093: 
1094: \subsection{Enhancements and Stochastic Acceleration}
1095: 
1096: In this section we describe the role the SA model described in the previous
1097: sections can play in the enrichment of \3he and other elements. The formalism
1098: described above can be applied to the acceleration of all ions in the same way
1099: used to fit the radiative signatures of the accelerated electrons and protons to
1100: observations. All the results presented below are based on acceleration by PPWs
1101: and most of the results are from Liu \ea (2004, 2006).
1102: 
1103: 
1104: 
1105: \begin{figure}[hbtp]
1106: \centerline{
1107: \includegraphics[height=7.5cm,angle=0]{ppwdispHe.eps}
1108: \includegraphics[height=7.5cm,angle=0]{timesHe.eps}
1109: }
1110: \caption
1111: {\scriptsize
1112: {\bf Left panel:} Dispersion relations for parallel propagating waves and
1113: resonance conditions for \3he and \he4 for the specified energies and three
1114: different values of $\alpha$.  Note that the resonance lines for \3he start (at
1115: $k$=0) at its gyrofrequency (divided by $\gamma\sim 1$), which is larger (in
1116: absolute value) than that of \he4. As a result at low energies \3he  has more
1117: resonant intersections (filled circles)  than \he4. {Right panel:} Acceleration,
1118: escape and energy time scales
1119: for \3he an \he4. The main difference here is the shorter acceleration time for
1120: \3he at lower energies because of its stronger interactions with waves.
1121: }
1122: \label{Hetimes}
1123: \vspace{0.5cm}
1124: \end{figure}
1125: 
1126: Figure \ref{Hetimes} (left) shows the dispersion relation near the proton and
1127: \he4 gyrofrequencies and the resonant conditions for two different energies of
1128: \3he and \he4 ions. The presence of fully ionized \he4 (about 8\% of protons) in
1129: the background plasma divides the Alfv\'en wave-proton cyclotron branch into two
1130: distinct modes which we have labeled proton cyclotron (PC) and He cyclotron
1131: (HeC) branches. Because of the gap between these two modes at low values of
1132: the wavevector $k$,  low energy \he4 ions resonate only with the HeC branch
1133: while
1134: \3he ions can resonate with both branches and are
1135: accelerated more efficiently. The right panel of this figure depicts the
1136: relevant
1137: time scales showing the much shorter acceleration time for \3he than \he4 at low
1138: energies. This is the primary cause of the \3he enrichment.
1139: 
1140: As a result of this more efficient acceleration most \3he ions are accelerated
1141: to form a non-thermal distribution. While \he4 ions are heated up forming a low
1142: energy quasi-thermal bump with only a small fraction reaching to high energies.
1143: Figure \ref{Hespectra} shows model spectra fitted to two \3he rich SEP events
1144: displaying the above features.  Another result of
1145: the efficient acceleration is that one expects a high degree of acceleration of
1146: \3he even in smaller events such that the \3he/\he4 abundance ratio is expected
1147: to decrease with increasing fluence of the event. This is what is seen (see Fig.
1148: XXX in Chapter 2) and is demonstrated in the right panel of Figure
1149: \ref{Hespectra} where we show spectra for three different values of the
1150: acceleration rate $\tau_p^{-1}$. For higher rates more of the \he4 ions in the
1151: quasi-thermal tail reach into the non-thermal tail decreasing the \3he
1152: enrichment and the \3he/\he4 ratio.
1153: 
1154: \begin{figure}[hbtp]
1155: \centerline{
1156: \includegraphics[width=5.8cm,angle=0]{he34Aug.eps}
1157: \hspace{-0.5cm}
1158: \includegraphics[width=5.8cm,angle=0]{spectra_taup.ps}
1159: \hspace{-0.5cm}
1160: \includegraphics[width=5.8cm,angle=0]{spectra_alpha.eps}
1161: }
1162: \caption
1163: {\scriptsize Fits to \3he and \he4 spectra of two \3he rich SEP events. Note
1164: that because of the shorter acceleration time of \3he at low energies most of
1165: the \3he is accelerated into super thermal tail while most of \he4 ions are
1166: heated up to a high 'temperature' quasi-thermal distribution with only a smaller
1167: fraction
1168: forming a non-thermal tail. The middle and right panels also show two additional
1169: sets of
1170: model spectra with somewhat higher and lower values of $\tau_p^{-1}$, which is
1171: proportional to the turbulence energy density or the rate of acceleration,  and
1172: $\alpha$, respectively.
1173: }
1174: \label{Hespectra}
1175: \vspace{0.5cm}
1176: \end{figure}
1177: 
1178: The dependence of \3he enrichment on model parameters is further 
1179: demonstrated in Figure \ref{Hetaup}. The most influential parameter is the
1180: acceleration rate $\tau_p^{-1}\propto {\cal W}$ and there is some variation with
1181: $\alpha$ at low
1182: acceleration rates. There is little dependence on the (unknown) spectral
1183: parameters of the turbulence. The trend and the range of the
1184: variations presented here are in agreement with the observations described in
1185: chapter 2. It remains to be seen whether or not the spectral trend presented in
1186: Figure \ref{Hespectra} (middle and right) also agrees with observations. It
1187: should also be possible
1188: to use the observed distributions of \he4 and \3he fluences or peak fluxes to
1189: further test this promising model and constrain its parameters. This will
1190: require a sample of events with known selection criteria and observational
1191: biases.
1192: 
1193: \begin{figure}[hbtp]
1194: \centerline{
1195: \includegraphics[height=7.0cm,angle=0]{ratio_taup.eps}
1196: \includegraphics[height=7.0cm,angle=0]{ratio_taupk.eps}
1197: }
1198: \caption
1199: {\scriptsize Variation with acceleration rate (or level of turbulence) of the
1200: \3he/\he4 fluence ratio at 2 MeV per nucleon for different values of $\alpha$
1201: (left) and the
1202: spectral break wave vector $k_\max$ (right). The black lines are for best fit
1203: values of $\alpha=0.46$ and  $k_\min=0.1 k_\max$ for the 30 Sep. 1999 event
1204: shown
1205: in Figure \ref{Hespectra} (middle and left). These curves  show the possibility
1206: of
1207: obtaining a large range of the ratio of \3he/\he4 and that the main
1208: characteristics of turbulence affecting this ratio is the energy density ${\cal
1209: W}\propto \tau_p^{-1}$ with smaller effect due to $\alpha$ and a relatively
1210: insignificant effect due to unknown details of the turbulence spectrum (see Liu
1211: \ea 2006).
1212: }
1213: \label{Hetaup}
1214: %\vspace{0.5cm}
1215: \end{figure}
1216: 
1217: \section{Summary and Discussion}
1218: 
1219: We have discussed possible mechanism of  accelerations in solar flares and SEPs,
1220: and pointed out that in all
1221: mechanisms plasma  waves and turbulence play a major role. We have then argued
1222: that in solar flares the most likely mechanism is stochastic acceleration  by 
1223: plasma  waves and turbulence and have demonstrated
1224: that the predictions of this model are consistent with most of the observed
1225: flare
1226: radiative signatures. We have also shown that the observed \3he enrichment and
1227: the spectra of \3he and \he4 can  be described by such a model without
1228: requiring any special conditions. The same model that successfully describes the
1229: observed radiation can also account for this long standing puzzle. Unlike
1230: previous models, which addressed only the enhancement (without detailed
1231: comparison with observed spectra) and required  special conditions and/or two
1232: stage acceleration, the  model presented here neither requires special
1233: conditions nor another acceleration mechanism.
1234: 
1235: Another important problem of SEP observations is enhancement of abundances of
1236: heavier ions with the enhancement increasing with ionic mass or mass-to-charge
1237: ratio described in Chapter 2. Here the situation is more complicated because of
1238: the
1239: uncertainty about the charge state of these ions and the observations of
1240: dependence of the charge state on the ion energy discussed in previous chapters.
1241: These observations also favor
1242: acceleration in relatively high densities of the lower corona, where the
1243: stochastic acceleration 
1244: invoked here occurs. Figure  \ref{heavies} (left)  shows abundances of a large
1245: number of ion species obtained from a
1246: composite of many SEP events (red points, Mason, private communication). The
1247: black points and the line show
1248: our preliminary attempt (Liu \& Petrosian in preparation) to fit these
1249: observations. Fits to protons \3he and
1250: \he4 are as above and in plotting the line we have assumed a charge of 16 for
1251: all the elements above iron.
1252: 
1253: \begin{figure}[hbtp]
1254: \centerline{
1255: \includegraphics[height=7.0cm,angle=0]{heavyionflux.ps}
1256: \includegraphics[height=7.0cm,angle=0]{oxygen.ps}
1257: }
1258: \caption
1259: {\scriptsize {\bf Left panel:} A preliminary fit by the SA model to different
1260: ion enhancements obtained from a composite of many events (red points from
1261: Mason, private communication). The open circles and the black curve are the
1262: model
1263: results assuming a charge of 16 for all elements heavier than iron (Liu \&
1264: Petrosian, in preparation). This, of course, is inappropriate for intermediate
1265: ions where the time-independent model described here cannot describe the
1266: observations. {\bf Right panel:} A preliminary result from a time-dependent
1267: treatment showing the evolution of $^{16}$O (dashed lines) spectrum and the time
1268: integrated spectrum (solid green) fitted to  observed spectrum of fluences of
1269: $^{16}$O and \he4 for a typical event.
1270: }
1271: \label{heavies}
1272: %\vspace{0.5cm}
1273: \end{figure}
1274: 
1275: This scenario in the form presented here will encounter difficulty with the
1276: observations of the intermediate ions (\eg CNO). We note that most of the
1277: observed spectra refer to the event fluence. The models described above are
1278: based on the \underline{time-independent} (or the steady state) equations. This
1279: is valid when the acceleration time is shorter than the dynamic time scale of
1280: the
1281: event (\eg event duration), in which case the steady state results are good
1282: approximation. A more accurate analysis requires a \underline{time-dependent}
1283: analysis. Figure \ref{heavies} (right) shows a preliminary result from such a
1284: model (East, Petrosian \& Liu, in preparation) from such a model. The dashed
1285: lines
1286: show the time evolution of the spectrum of $^{16}$O and the solid green line the
1287: time integrated spectrum appropriate for the fluence spectrum shown by the green
1288: squares. This model seems to provide a reasonable fit (except at low energies
1289: where the uncertainties are large) for both  $^{16}$O  and \he4. 
1290: Clearly more data and more detailed model are
1291: required to put this comparison on the same footing as those described above for
1292: electrons, protons and He ions.
1293: 
1294: \section{REFERENCES}
1295: 
1296: %\vspace{0.4cm}
1297: %\setlength{\baselineskip}{8pt}
1298: %{\footnotesize
1299: %\noindent
1300: \def\et{{\it et~al.}}
1301: \def\aa{{\it Astr. \& Astrophys.}}
1302: \def\anyas{{\it Ann. N.Y. Acad. Sci.}}
1303: \def\apj{{\it Ap.~J.},\,\,}
1304: \def\apjl{{\it Ap.~J. (Letters),\,\,}}
1305: \def\ass{{\it Ap. Space Sci.}}
1306: \def\jgr{{\it J.~Geophys. Res.}}
1307: \def\mnras{{\it M.N.R.A.S.}}
1308: \def\nature{{\it Nature}}
1309: \def\prd{{\it Phys. Rev.~D}}
1310: \def\pra{{\it Phys. Rev.~A}}
1311: \def\ssr{{\it Space Sci. Rev.}}
1312: \def\rmp{{\it Rev. Mod. Phys.}}
1313: \def\sphys{{\it Solar Phys.}}
1314: \def\JL{Leach,~J.\ }
1315: \def\VP{Petrosian,~V.\ }
1316: 
1317: 
1318: 
1319: \def\refer{ \par \noindent \hangindent=2pc \hangafter=1}
1320: \baselineskip = 10 true pt
1321: %\bigskip
1322: 
1323: \refer Amato, E., \& Blasi, P. 2005, MNRAS, 364, 76
1324: \refer Andr\'e, Mats 1985, {\it J. Plasma Phys.,} 33, 1
1325: \refer Boris, J. P., Dawson, J. M., Orens, J. H., \& Roberts, K. V. 1970, 
1326: {\it Phys. Rev. Lett.}, 25, 706
1327: \refer Cassak, P. A., Drake, J. F. \& Shay, M. A. 2006, \apj 644, L145
1328: \refer Chandran, B. 2005, {\it Phys. Rev. Lett.}, 95, 265004
1329: \refer Cho, J., \& Lazarian, A. 2002, {\it Phys. Rev. Lett.}, 88, 245001
1330: \refer Cho, J., \& Lazarian, A. 2006, {\it Particle Acceleration by MHD Turbulence}, 
1331: \apj 638, 811
1332: \refer Chupp, E. L. 1990, {\it Science}, 250, 229
1333: \refer Cranmer, S. R., \& van Ballegooijen, A. A. 2003, \apj 594, 573
1334: \refer Dennis, B. R., \& Zarro, D. M. 1993, \sphys, 146, 177
1335: \refer Dingus, B. L. et al. 1994, {\it AIP Conf. Proc.} eds. J. M. Ryan \& W. T. Vestrand, 294, 177
1336: \refer Drake, J. F., Paper presented at {\it Krakow Conference on Relativistic Jets}, June  2006
1337: \refer Ellison, D. C., Decourchelle, A, \& Ballet, J. 2005, \aa,  429, 569
1338: \refer Farmer, A., Goldreich, P. 2004, \apj 604, 671
1339: \refer Fisk, L. A. 1978, \apj 224, 1048
1340: \refer Giacalone, J. 2005, \apjl 628, L37
1341: \refer Goldreich, P., \& Sridhar, S. 1995, \apj 438, 763
1342: \refer Goldreich, P., \& Sridhar, S. 1997, \apj 485, 680
1343: \refer Hamilton, R. J. \& Petrosian, V. 1992, \apj 398, 350
1344: \refer Holman, G. D. 1985, \apj {\bf 293}, 584
1345: \refer Holman, G. D. 1996a, {\it BAAS}, 28, 939
1346: \refer Holman, G. D. 1996b,  {\it AIP Conf. Proc.} eds. R. Ramaty et al., 374, 479
1347: \refer Hoshino, M., Paper presented at {\it COSPAR Scientific Assembly}  Beijing, July 2006
1348: \refer Hurford, G. J. et al. 2003, \apj 595, 77L
1349: \refer Ibragimov, I. A. \& Kocharov, G. E. 1977, Proc. 15th Int, Cosmic Ray Conf. (Plovdiv), 11,  340
1350: \refer Iroshnikov, P., 1963, {\it Astron. Zh.}, 40, 742 (English version: 1964,  Sov. Astron., 7, 566)
1351: \refer Jiang, Y. W. et al. 2006,  \apj 638, 1140 
1352: \refer Jones, F. C. 1994, {\it Ap. J. Suppl.}, 90, 561
1353: \refer Jokipii, R. J. 1987, \apj 313, 842
1354: \refer Kraichnan, R., 1965, Phys. Fluids, 8, 1385
1355: \refer Lin, R. P. et al. 1981, \apjl 251, L109
1356: \refer Lithwick, Y., \& Goldreich, P. 2001, \apj 562, 279 
1357: \refer Litvinenko, Y. E. 2003, {\it Solar Phys.}, 212, 379
1358: \refer Liu, S., Petrosian, V., \& Mason, G. 2004, \apjl 613, L81 
1359: \refer Liu, S., Petrosian, V., \& Mason, G. 2006, \apj 634, 462 
1360: \refer Liu, W. et al. 2004, \apj 611, 53L
1361: \refer Liu, W.  2007, from PhD thesis, Stanford University
1362: \refer Luo, Q. \& Melrose, D. 2006; astro-ph/0602295
1363: \refer Kulsrud, R.M. 2005, {\it Plasma Physics for Astrophysics}, Princeton Univ. Press
1364: \refer Mandzhavidze, N., \& Ramaty, R. 1996, {\it BAAS}, 28, 858
1365: \refer Marschh\"{a}user, H. et al. 1994, {\it AIP Conf. Proc.} eds. J. M. Ryan \& W. T. Vestrand 294, 171
1366: \refer Mason, G. M. et al. 1986, \apj 303, 849
1367: \refer Mason, G. M., Dwyer, J. R., \& Mazur, J. E. 2000, \apj 545, 157L
1368: \refer Mason, G. M. et al. 2002, \apj 574, 1039
1369: %\refer Mason, G. M. et al. 2004, \apj 605, 555
1370: \refer Masuda, S. et al. 1994, {\it Nature}, 371, 495
1371: \refer Masuda, S. 1994, University of Tokyo, Ph.D. Thesis
1372: \refer Mazur, J. E. et al. 1992, \apj 401, 398
1373: \refer McTiernan, J. M. et al. 1993, \apjl 416, L91
1374: \refer Miller, J. A.\ 2003, {\it COSPAR Colloquia Series} Vol. 13, 387
1375: \refer Miller, J. A.,LaRosa, T.N., \& Moore, R.L. 1996, \apj 445, 464
1376: \refer Miller, J. M. \&  Reames, D. V. 1996, {\it AIP Conf. Proc.} eds. R. Ramaty et al., 374, 450
1377: \refer Miller, J.A.\& Vi\~vas, A. F. 1993, ApJ, 412, 386
1378: \refer Neupert, W.M. 1968, \apj 153, L59
1379: \refer  Park, B. T., Petrosian, V., \& Schwartz, R. A. 1997, \apj 489, 358
1380: \refer Paesold, G., Kallenbach, R. \& Benz, A. O. 2003, ApJ 582, 495 
1381: \refer Petrosian, V. 1973, \apj 186, 291
1382: \refer Petrosian, V., \& Donaghy, T. Q. 1999, \apj 527, 945 
1383: \refer Petrosian, V.,  Donaghy, T. Q., \& McTiernan, J. M. 2002, \apj 569, 459 
1384: \refer Petrosian, V., \& Liu, S. 2004, \apj 610, 550 {\bf PL04}
1385: \refer Petrosian, V., McTiernan, J.M., \& Marschh\"{a}user, H. 1994, \apj 434, 744
1386: \refer Petrosian, V,, Yan, H. \& Lazarian, A. 2006, \apj 644, 603 
1387: \refer Pryadko, J., \& Petrosian, V. 1997, \apj 482, 774 
1388: \refer Pryadko, J., \& Petrosian, V. 1998, \apj 495, 377 
1389: \refer Pryadko, J., \& Petrosian, V. 1999, \apj 515, 873 
1390: \refer Ramaty, R. \& Kozlovsky, B. 1974, ApJ 193, 729
1391: \refer Reames, D. V., Meyer, J. P., \& von Rosenvinge, T. T.\ 1994, ApJS, 90, 649
1392: \refer Reames, D. V. et al. 1997, \apj 483, 515
1393: %\refer R\"onnmark, K. 1983, {\it Plasma Phys.}, 25, 669
1394: \refer Sui, L., \& Holman, G. D. 2003, \apjl 596, L251
1395: \refer Sui, L., Holman, G. D., \& Dennis, B. R. 2005, \apj 626, 1102
1396: \refer Swanson, D. G. 1989, {\it Plasma Waves} (New York: Academic Press)
1397: \refer Temerin, M. \&  Roth, I. 1992, ApJ 391, L105
1398: \refer Tsuneta, S. 1985, \apj 290, 353
1399: \refer Veronig, A. M. et al. 2005, \apj 621, 482
1400: \refer Yan, H., \& Lazarian, A. 2002, {\it Phys. Rev. Lett.}, 89, 2881102
1401: \refer Zenitani, S. \& Hoshino, M. 2005, \apj 618, L111
1402: \refer Zhang, T. X. 1995, ApJ, 449, 916
1403: \refer Zhou, Y., \& Matthaeus, W.H. 1990, {\it J. Geophys. Res.} 95, 14881
1404: 
1405: 
1406: \end{document}
1407: 
1408: 
1409: 
1410: