-
Notifications
You must be signed in to change notification settings - Fork 0
/
chap2.tex
588 lines (551 loc) · 43.3 KB
/
chap2.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
%
% This is Chapter 2 file (chap2.tex)
%
\chapter{Kinetic theory and linear microkinetic instabilities}\label{chap:chap2}
Kinetic theory forms a significant part of our understanding of plasma, specially at small
scales (less than a $d_{\rm i}$). In this chapter, we give a brief overview of the salient
properties of this theory. In \Cref{sec:intr2} we start with the equation of motion for a
charged particle in an electromagnetic field and extend this idea to an ensemble of particles.
In \Cref{sec:instab} we discuss anisotropy and linear instabilities arising because of it. We
also discuss how one can compute rate of growth of these instabilities using kinetic theory. In
\Cref{sec:app2} we discuss some of the application and observational evidence of linear theory.
We finish this chapter with a brief discussion of limitations of linear theory in
\Cref{sec:conc2}.
\section{Introduction to Plasma Kinetic Theory} \label{sec:intr2}
\subsection{Equations of motion}\label{sec:eqom}
Consider a system which consists of a single charged particle of mass $m$ and charge $q$
in a magnetic field $\mathbf{B}$ and an electric field $\mathbf{E}$. The
non-relativistic equation of motion for this particle can then be written as:
\begin{align}
m \frac{d\mathbf{v}_1}{dt} & = q\left(\mathbf{E} + \mathbf{v}_1 \times \mathbf{B}\right) \label{eq:emot1}
\end{align}
and the its exact phase space\index{phase space} density at any point in space can be written as :
\begin{align}
\mathcal{F}_1(\mathbf{x},\mathbf{v},t) & = \delta \left(\mathbf{x}-\mathbf{x}_1(t)\right)\delta \left(\mathbf{v}-\mathbf{v}_1(t)\right) \label{eq:dens1}
\end{align}
where $\mathbf{x}_1(t)$ and $\mathbf{v}_1(t)$ are the position and velocity of the
particle at any time $t$ and $\delta(...)$ is the Dirac delta function. The six
dimensional space spanned by $\mathbf{x}$ and $\mathbf{v}$ is called phase
space\footnote{Note: $\mathbf{x}$ and $\mathbf{v}$ are independent coordinates in phase
space.}. If we have $n$ such particles in the system, then for the $i^\mathrm{th}$
particle, \Cref{eq:emot1} can be written as :
\begin{align}
m_{\rm i} \frac{d\mathbf{v}_{\rm i}}{dt} & = q_{\rm i}\left(\mathbf{E}_\mu + \mathbf{v}_{\rm i} \times \mathbf{B}_\mu\right) \label{eq:emotn}
\end{align}
where the subscript $\mu$ represents the superposition of all the fields exerted by the
particles in the system at the position of $i^\mathrm{th}$ particle. The total phase
space density can now be written as:
\begin{align}
\mathcal{F}_{\rm n}(\mathbf{x},\mathbf{v},t) & = \sum_{i=1}^{n}\delta \left(\mathbf{x}-\mathbf{x}_{\rm i}(t)\right)\delta \left(\mathbf{v}-\mathbf{v}_{\rm i}(t)\right) \label{eq:densn}
\end{align}
In a closed system where there is no addition or removal of any particle, the total
phase space density of a fluid element in phase space will remain constant in time.
Using this conservation of phase space density one can write:
\begin{align}
\frac{d}{dt}\left(\mathcal{F}_{\rm n}(\mathbf{x},\mathbf{v},t)\right) & = 0 \label{eq:cons}
\end{align}
Since both $\mathbf{x}$ and $\mathbf{v}$ depend on time, we can use the chain rule and
write \Cref{eq:cons} as:
\begin{align}
\frac{\partial \mathcal{F}_{\rm n}}{\partial t} + \frac{d \mathbf{x}}{d t} \cdot \frac{d \mathcal{F}_{\rm n}}{d \mathbf{x}} + \frac{d \mathbf{v}}{d t} \cdot \frac{d \mathcal{F}_{\rm n}}{d \mathbf{v}} & = 0 \label{eq:cons2}
\end{align}
where $\frac{\partial}{\partial t}$ is the partial derivative with respect to $t$. In
\Cref{eq:cons2} we have dropped $(\mathbf{x},\mathbf{v},t)$ for the sake of readability.
Using the fact that $\frac{d}{d t}\mathbf{x}=\mathbf{v}$ and substituting for $\frac{d
\mathbf{v}}{d t}$ from \Cref{eq:emotn} in \Cref{eq:cons2}, we have:
\begin{align}
\frac{\partial \mathcal{F}}{\partial t} + \mathbf{v} \cdot \nabla_{\mathbf{x}} \mathcal{F} + \frac{q}{m}\left(\mathbf{E}_\mu + \mathbf{v} \times \mathbf{B}_\mu\right) \cdot \nabla_{\mathbf{v}} \mathcal{F} & = 0 \label{eq:cons3}
\end{align}
Solving \Cref{eq:cons3} (called the Klimontovich-Dupree equation) is quite a difficult
task since it contains all the microscopic fields, the computation of which involves
tracking the position and velocity of all the particles, which as we have already
discussed is quite impossible to implement. This problem can be mitigated by writing the
phase density as a sum of two parts, average and fluctuating as below:
\begin{align}
\begin{split}
\mathcal{F} & = \left<\mathcal{F}\right> + \delta\mathcal{F} \label{eq:pd1}\\
& = \mathnormal{f} + \delta\mathcal{F}
\end{split}
\end{align}
where $\left<\mathcal{F}\right>$ denotes the smoothed average or the background value of
$\mathcal{F}$, and $\delta \mathcal{F}$ is the fluctuation in the smoothed
$\mathcal{F}$. For ease of writing, from now on $\left<\mathcal{F}\right>$ will simply
be denoted by $\mathnormal{f}$, which is also called the distribution function\index{distribution function} and is
interpreted as the probability of finding a particle at any location within a phase
space volume d\textbf{x}d\textbf{v}. If we carry out the same process for the fields, we
can write them as:
\begin{align}
\begin{split}
\mathbf{E}_{\mu} & = \left<\mathbf{E}_{\mu}\right> + \delta\mathbf{E}_{\mu} \\
& = \mathbf{E} + \delta\mathbf{E}_{\mu}
\end{split}
\begin{split}
\mathbf{B}_{\mu} & = \left<\mathbf{B}_{\mu}\right> + \delta\mathbf{B}_{\mu} \\
& = \mathbf{B} + \delta\mathbf{B}_{\mu}\label{eq:pd2}
\end{split}
\end{align}
Since $\mathcal{F}$, \textbf{B} and \textbf{E} are all smoothed averages, their
fluctuations $(\delta\mathcal{F}, \delta\mathbf{B}, \delta\mathbf{E}$) will form an
statistical ensemble which would imply that $\left<\delta ... \right> = 0$.
Using \Cref{eq:pd1,eq:pd2} in \Cref{eq:cons3} and then taking the ensemble average, we
have:
\begin{align}
\frac{\partial \mathnormal{f}}{\partial t} + \mathbf{v} \cdot \nabla_{\mathbf{x}}\mathnormal{f} + \frac{q}{m}\left(\mathbf{E} + \mathbf{v} \times \mathbf{B}\right) \cdot \nabla_{\mathbf{v}} \mathnormal{f} & = \frac{q}{m}\left<\left(\delta\mathbf{E} + \mathbf{v} \times \delta \mathbf{B}\right) \cdot \nabla_{\mathbf{v}} \mathcal{F}\right> \label{eq:kint}
\end{align}
This is the kinetic equation and defines the evolution of phase space density with time
and position.
Computing the right hand side of \Cref{eq:kint} is quite a difficult task, thus we often
assume that the correlation between the background field and its fluctuation is
infinitely small and collisions between particles account for correlation among
themselves and occur uncorrelated of each other as random events, a consequence of the
molecular chaos hypothesis\index{chaos hypothesis} (or
\discolorlinks{\href{https://en.wikipedia.org/wiki/Boltzmann\_equation\#The\_collision\_term\_(Stosszahlansatz)\_and_molecular\_chaos}{\textit{sto\ss
zahlansatz}}}) \citep{ClerkMaxwell1867}. Under these assumptions we can simply replace
the right side of \Cref{eq:kint} with a collision operator\index{collision operator}, $\left(\frac{\partial
\mathnormal{f}}{\partial t}\right)_{\rm c}$, which ideally will have all the information
related to particle-particle interaction. \Cref{eq:kint} can thus be re-written as:
\begin{align}
\frac{\partial \mathnormal{f}}{\partial t} + \mathbf{v} \cdot \nabla_{\mathbf{x}}\mathnormal{f} + \frac{q}{m}\left(\mathbf{E} + \mathbf{v} \times \mathbf{B}\right) \cdot \nabla_{\mathbf{v}} \mathnormal{f} & = \left(\frac{\partial \mathnormal{f}}{\partial t}\right)_{\rm c} \label{eq:bolt}
\end{align}
This is the well known Boltzmann equation from statistical mechanics.
\citet{Baumjohann1996} and references therein give some details about different
functional forms of collisional operators\index{collision operator}.
Presence of collision can significantly alter the shape of a VDF (see \Cref{sec:dfd} for
definition). Since collisions can often result in transfer or exchange of energy and
momentum between particles they work to erode the non-equilibrium features of a VDF
(more on this later). In a fully ionized plasma where collisions are primarily coulombic
in nature, computing the collision operator is further complicated by its dependence on
temperature and density \citep{Baumjohann1996}. \citet{Landau1936} computed the
collision operator, $\left(\frac{\partial \mathnormal{f}}{\partial t}\right)_{\rm c}$,
for such a plasma, though solving it even for the simplest of cases is quite a daunting
ask \citep{Verscharen2019}. However, \citet{Marsch1982, Marsch2010} showed that for
space plasmas in the inner heliosphere collisions play negligible to minor role. We thus
simply set the value of collision operator to zero and the Boltzmann equation
(\Cref{eq:bolt}), in the absence of collision reduces to the Vlasov\index{Vlasov!equation} equation:
\begin{align}
\frac{\partial \mathnormal{f}}{\partial t} + \mathbf{v} \cdot \nabla_{\mathbf{x}}\mathnormal{f} + \frac{q}{m}\left(\mathbf{E} + \mathbf{v} \times \mathbf{B}\right) \cdot \nabla_{\mathbf{v}} \mathnormal{f} & = 0 \label{eq:vlas}
\end{align}
This equation forms the basis of much of kinetic theory for space plasmas and is used
extensively in this thesis. We also use the Vlasov equation to derive the dispersion
relation (see\,\Cref{sec:dpr}), which gives us an idea about the kind of waves and
instabilities present in the system (more on this later).
\subsection{Distribution Function and Other Definitions} \label{sec:dfd}
As discussed in \Cref{sec:eqom}, $\mathnormal{f}(\mathbf{x,v},t)$ gives the probability
density function (PDF)\index{distribution function!PDF} of particles in phase space. Different statistical moments of the
PDF gives us macroscopic properties of the whole ensemble. Often, we are interested in
the behaviour of the system at one particular position in the configuration space at any
given time. This would mean that the PDF will have no dependence on the position
(\textbf{x}) or any explicit dependence on time (t). We thus have a PDF with explicit
dependence on just the velocity given as $\mathnormal{f}(\mathbf{v})$. This is referred
to as velocity distribution function (VDF)\index{distribution function!VDF}.
For a VDF we can define some parameters associated with plasma using its statistical
moments.\\
\textbf{Density ($n_{\rm j}$)}: Number density of species $j$, usually expressed in the
units of $\mathrm{cm^{-3}}$ and can be derived from VDF by computing its zeroth order
moment as follows:
\begin{align}
n_{\rm j} = \int_{\forall \mathbf{v}} d^3\mathrm{v} \mathnormal{f}_{\rm j}\left(\mathbf{v}\right) \label{eq:numden}
\end{align}
\textbf{Bulk velocity\index{Bulk velocity} ($\mathbf{v}_{\rm j}$)}: The bulk velocity of the species $j$, or
the mean particle velocity, usually has units of $\mathrm{km/sec}$ and can be derived
from the VDF's first-order moment:
\begin{align}
\mathbf{v}_{\rm j} = \frac{1}{n_{\rm j}}\int_{\forall \mathbf{v}} d^3\mathrm{v} \mathbf{v} \mathnormal{f}_{\rm j}\left(\mathbf{v}\right) \label{eq:blkvel}
\end{align}
\textbf{Thermal Speed\index{Thermal Speed} ($w_{\rm j}$)}: It represents the thermal speed of the species j
and has units of $\mathrm{km/sec}$. It is a measure of thermal energy of species $j$ and
the value for it can be derived from the VDF's second-order moment:
\begin{align}
\mathrm{v}_{\rm j}^2 + 3\,w_{\rm j}^2 = \frac{1}{n_{\rm j}}\int_{\forall \mathbf{v}} d^3\mathrm{v} \mathbf{v}^2 \mathnormal{f}_{\rm j}\left(\mathbf{v}\right) \label{eq:wtherm}
\end{align}
For an anisotropic case, where temperature is different along different directions,
\Cref{eq:wtherm} is computed for each component separately (see \citet[\S
1.4.1]{Verscharen2019} for a detailed description).\\
\\
\textbf{Temperature ($T_{\rm j}$)}: Using the thermal speed, one can define the
temperature (in units of K) of the species as:
\begin{align}
T_{\rm j} & = \frac{m_{\rm j}\,w_{\rm j}^2}{2k_{\rm B}} \label{eq:temp}
\end{align}
In a magnetized plasma, the VDFs commonly exhibit distinct temperatures perpendicular
and parallel to the magnetic field because of the slightly different heating or cooling
rates in different directions \citep{Stix1992, Gary1993}. We thus have different
temperatures in parallel ($T_{\rm \parallel j}$) and perpendicular ($T_{\rm \perp j}$)
directions and the total temperature of species j is then given as:
\begin{align}
T_{\rm j} = \frac{T_{\rm \parallel j} + 2 T_{\rm \perp j}}{3} \label{eq:ttemp}
\end{align}
The ratio of two temperatures (perpendicular and parallel) is called anisotropy\index{anisotropy} and is
expressed as:
\begin{align}
R_{\rm j} = \frac{T_{\rm \rm \perp j}}{T_{\rm \parallel j}} \label{eq:aniso}
\end{align}
\textbf{Parallel Beta ($\beta_{\rm \parallel j}$)}: It is the ratio of parallel thermal
energy of a species to the magnetic pressure energy stored in the field.
\begin{align}
\beta_{\rm \parallel j} = \frac{n_{\rm j}\,k_{\rm B}\,T_{\rm \parallel j}}{B_0^2/(2\mu_\circ)} \label{eq:beta}
\end{align}
As we will see later in this chapter (see \Cref{sec:instab2}) the values of parameters
$R_{\rm j}$ and $\beta_{\rm \parallel j}$ play an important role in determining if a
region of plasma is stable or unstable. \\
\\
\textbf{Shape of a VDF}:\\
In general the VDF can take a variety of forms as long as they conform to the laws of
probability. For a plasma in local thermodynamic equilibrium, it takes the shape of a
Maxwellian distribution\index{distribution function!Maxwellian} as shown in \Cref{eq:vdf}.
\begin{align}
\mathnormal{f}_{\rm j}(\mathbf{v}) & = \frac{n_{\rm j}}{\left(\pi w_{\rm j}^2\right)^{3/2}} \mathrm{exp}\left(-\frac{\left|\mathbf{v} - \mathbf{v}_{\rm 0}\right|^2}{w_{\rm j}^2}\right) \label{eq:vdf}
\end{align}
where $\mathbf{v}_{\rm 0}$ is the streaming velocity of the plasma. Each species ($j$)
in the plasma has its own VDF, the statistical moments of which give the species' bulk
parameters (e.g., density and velocity).
One can define a direction based on the background field present in the plasma. If we
assign the direction along the magnetic field as the parallel direction (represented by
$\parallel$) and the other two orthogonal directions as the perpendicular direction
(represented by $\perp$), the total magnetic field can be expressed in this new
coordinate system as:
\begin{align}
\mathbf{B} = B_\parallel\,\mathbf{\hat{e}}_\parallel + B_\perp\,\mathbf{\hat{e}}_\perp \label{eq:newdir}
\end{align}
where $\mathbf{\hat{e}}_\parallel~\mathrm{and}~\mathbf{\hat{e}}_\perp$ are the unit
vectors along and perpendicular to the magnetic field.
It is often easier to work with VDFs in this coordinate system, thus we rewrite
\Cref{eq:vdf} for a species $j$ in the new coordinate system as:
\begin{align}
\mathnormal{f}_{\rm j}(\mathbf{v}) & = \frac{n}{\pi^{3/2} w_{\rm \perp j}^2 w_{\rm \parallel j}} \mathrm{exp}\left(-\frac{(\mathrm{v}_\parallel - \mathrm{v}_{\rm \parallel 0j})^2}{w_{\rm \parallel j}^2} -\frac{|\mathbf{v}_\perp - \mathbf{v}_{\rm \perp 0j}|^2}{w_{\rm \perp j}^2}\right) \label{eq:vdf2}
\end{align}
The values of $\mathbf{v}_{\rm \parallel 0j}$ and $\mathbf{v}_{\rm \perp 0j}$ are often
different resulting in a slightly different VDF in the two directions. For such a case,
the VDF is referred to as the bi-Maxwellian\index{distribution function!bi-Maxwellian} distribution.
Though in this thesis we use \Cref{eq:vdf2} as the standard/default VDF unless otherwise
stated, it must be noted that for solar wind, especially at 1\,au, the VDF departs
significantly from a simple bi-Maxwellian \citep{Feldman1974, Feldman1974a, Marsch1982b,
Alterman2018}. Ion VDFs often have an asymmetry which can be more accurately accounted
for by superposition of a differentially flowing bi-Maxwellian \citep{Alterman2018}.
Other forms of distribution such as kappa distribution has also been used to study the
non-Maxwellian features of VDF like enhanced tail \citep{Pierrard2010,Pierrard2014,
Maksimovic1997,Nicolaou2020}.
\section{Linear Microkinetic Instabilities} \label{sec:instab}
\subsection{The Linear Dispersion Relation} \label{sec:dpr}
Though in \Cref{sec:eqom} we made several assumptions and used our a priori knowledge of
the system, solving \Cref{eq:vlas} for even a simple distribution like the bi-Maxwellian
(\Cref{eq:vdf2}) is quite complicated and computationally expensive. Coupling between
fields produced by one species with another species complicates it further.
Linearization (or linear analysis), where one assumes plain wave perturbation in fields
and the VDF helps simplify the problem while keeping the underlying physics of the
equations intact as long as the fluctuations have small amplitude relative to the
background values. In standard linear theory, we assume an equilibrium (i.e., constant)
background and perturb it with a small-amplitude sinusoidal fluctuation of wave vector
\textbf{k} and angular frequency $\omega$. The goal then is to derive, for a given
plasma, the dispersion relation: the relationship between \textbf{k} and $\omega$. Under
this assumption one can rewrite \Cref{eq:pd1,eq:pd2} as:
\begin{align}
\begin{split}
\mathnormal{f}_{\rm j}(\mathbf{x}, \mathbf{v}, t) & = \mathnormal{f}_{\rm j}^0(\mathbf{x}, \mathbf{v}) + \mathnormal{f}_{\rm j}^1(\mathbf{x}, \mathbf{v}, t)\\
& = \mathnormal{f}_{\rm j}^0(\mathbf{x}, \mathbf{v}) + \mathnormal{f}_{\rm j}^1(\mathbf{k}, \omega, \mathbf{v})~e^{(i(\mathbf{k}\cdot \mathbf{x} -\omega t))}\label{eq:vdfn1}
\end{split}
\end{align}
\vspace{-0.4cm}
\begin{align}
\begin{split}
\mathbf{B}(\mathbf{x}, t) & = \mathbf{B}^0(\mathbf{x}) + \mathbf{B}^1(\mathbf{x}, t)\\
& = \mathbf{B}^0(\mathbf{x}) + \mathbf{B}^1(\mathbf{k},\omega)~e^{(i(\mathbf{k}\cdot \mathbf{x} -\omega t))}\label{eq:mag1}
\end{split}
\end{align}
\vspace{-0.4cm}
\begin{align}
\begin{split}
\mathbf{E}(\mathbf{x}, t) & = \mathbf{E}^0(\mathbf{x}) + \mathbf{E}^1(\mathbf{x}, t)\\
& = \mathbf{E}^0(\mathbf{x}) + \mathbf{E}^1(\mathbf{k},\omega)~e^{(i(\mathbf{k}\cdot \mathbf{x} -\omega t))}\label{eq:efl1}
\end{split}
\end{align}
Where $\mathbf{k}$ and $\omega$ are the wavenumber vector and the frequency of
perturbation, respectively. \Crefrange{eq:vdfn1}{eq:efl1} along with Maxwell's
equation's (\Crefrange{eq:maxwell1}{eq:maxwell4}) help us drive the dispersion
relations. In order to do so, we start with some simple assumptions and conditions. We
work in a frame of reference where the zeroth order current density
($\mathbf{J}^0(\mathbf{x})$) and electric field ($\mathbf{E}^0(\mathbf{x})$) are zero
and the magnetic field ($\mathbf{B}^0(\mathbf{x})$) is constant. Thus
\Cref{eq:mag1,eq:efl1} can be rewritten as:
\begin{align}
\mathbf{B}(\mathbf{x}, t) & = \mathbf{B}_{\rm 0} + \mathbf{B}^1(\mathbf{x}, t)\label{eq:mag2}\\
\mathbf{E}(\mathbf{x}, t) & = \mathbf{E}^1(\mathbf{x}, t)\label{eq:efl2}\\
\mathbf{J}(\mathbf{x}, t) & = \mathbf{J}^1(\mathbf{x}, t)\label{eq:curr2}
\end{align}
We choose a co\"ordinate system with $z$-axis along the background magnetic field
($\mathbf{B}_{\rm 0}$). The angle between the direction of propagation of fluctuation
($\mathbf{k}$) and the magnetic field is:
\begin{align}
\cos \left(\theta \right) & = \frac{\mathbf{k}\cdot\mathbf{B}}{k B}\label{eq:theta}
\end{align}
Substituting expressions for $\mathbf{J}, \mathbf{B}~\mathrm{and}~\mathbf{E}$ into
Maxwell's equations (\Crefrange{eq:maxwell1}{eq:maxwell4}) and simplifying we have:
%by substituting $\frac{\partial}{\partial t}$ with $\omega$ and $\nabla$ with
%$\mathbf{k}$, we get:
\begin{align}
\mu_\circ\,\mathbf{J}^1(\mathbf{k}, \omega) & = \frac{i}{\omega}\,\mathbf{k} \times \left[\mathbf{k} \times \mathbf{E}^1(\mathbf{k}, \omega)\right] + \frac{i\,\omega}{c^2}\,\mathbf{E}^1(\mathbf{k}, \omega) \label{eq:curr3}
\end{align}
In similar fashion to the particle velocity in \Cref{eq:blkvel} one can write the flux
density as:
\begin{align}
\Gamma_{\rm j}^1(\mathbf{k}, \omega) = \int_{-\infty}^{\infty}d^3 v\,\mathbf{v}\,\mathnormal{f}_{\rm j}^1(\mathbf{k}, \omega, \mathbf{v}) \label{eq:flux1}
\end{align}
and thus define the current density as:
\begin{align}
\mathbf{J}^1(\mathbf{k}, \omega) & = \sum_{\rm j} q_{\rm j}\,\mathbf{\Gamma}_{\rm j}^1(\mathbf{k}, \omega) \label{eq:curr4}
\end{align}
For species $j$ we can define the dimensionless conductivity tensor $\mathbf{S}_{\rm j}$
as:
\begin{align}
\mathbf{\Gamma}_{\rm j}^1(\mathbf{k}, \omega) & = - \frac{i\,\epsilon_\circ\,k^2\,c^2}{q_{\rm j}\,\omega}\,\mathbf{S}_{\rm j}(\mathbf{k},\omega)\cdot \mathbf{E}^1(\mathbf{k},\omega) \label{eq:curr5}
\end{align}
Combining \Cref{eq:curr2,eq:curr3,eq:curr4} gives us:
\begin{align}
\mathbf{D}(\mathbf{k},\omega)\cdot \mathbf{E}^1(\mathbf{k},\omega) & = 0 \label{eq:disp1}
\end{align}
where,
\begin{align}
\mathbf{D}(\mathbf{k},\omega) = \left(\omega^2 - c^2\,k^2\right)\mathbf{I} + c^2\,\mathbf{k}\,\mathbf{k} + c^2\,k^2\,\sum_{\rm j}\,\mathbf{S}_{\rm j}(\mathbf{k},\omega) \label{eq:disp2}
\end{align}
with \textbf{I} being the 3-dimensional identity matrix and \textbf{k}\,\textbf{k} being
the dyadic\footnote{The dyadic product between two vectors \textbf{a} and \textbf{b} is
simply the product between the vector and its transpose, so one has
$\mathbf{a}\,\mathbf{b} = \mathbf{a}\,\mathbf{b}^T$, which results in a rank-two
tensor.} product of the wavevector. For \Cref{eq:disp2} to be true, $\mathbf{E}^1$
cannot be allowed to go to zero since that would allow perturbations in the system to
vanish. Hence we must have the determinant of $\mathbf{D}(\mathbf{k},\omega)$ go to
zero, thus:
\begin{align}
\mathrm{det}(\mathbf{D}(\mathbf{k},\omega)) & = 0 \label{eq:disp3}
\end{align}
which is the plasma dispersion relation\index{dispersion relation}. \citet{Gary1993} notes that this equation can
be solved ``either as a boundary value problem ($\omega$ is given as real, one solves
for a complex component of $\mathbf{k}$) or as an initial value problem ($\mathbf{k}$ is
given as real, one solves for complex $\omega$)''.
\Cref{eq:disp3} in general gives infinite number of solutions for $\omega$ for any value
of \textbf{k} and a set of plasma parameters. Thus $\omega (\mathbf{k})$ is a multi
valued function with multiple branches corresponding to different modes (see
\citet[Figure 2]{Schwartz1980}). However most modes are strongly damped and thus
dissipate before they can substantially effect the plasma.
\subsection{Temperature Anisotropy Induced Instabilities} \label{sec:instab2}
In a homogeneous plasma (PDF is independent of position) in local thermodynamic
equilibrium (LTE), all fluctuations (at any \textbf{k} and $\omega$) are damped and thus
have a decaying amplitude. However, departure from LTE because of the non-Maxwellian
properties of VDFs (temperature anisotropy or relative drift between two different
species, etc.) introduce free energy into the system. This results in circumstances
where instead of getting damped some perturbations grow exponentially and make the
system unstable. The rate at which such a perturbation propagates and gets damped or
grows is computed by solving the dispersion relation (\Cref{eq:disp3}). Solutions of
\Cref{eq:disp3} using the initial value problem method gives $\omega$, which is in
general complex and can be written as:
\begin{align}
\omega = \omega_{\rm r} + i \gamma \label{eq:omega}
\end{align}
where, $\omega_{\rm r}$ is the real component and $\gamma$ is the damping rate ($\gamma
< 0$) or growth rate\index{microinstability!growth rate} ($\gamma > 0$). In the presence of a growth rate, the fluctuations
present in the system start to grow exponentially. If the process continues,
fluctuations start having amplitudes comparable to the background value, which makes the
assumption of linearity void and makes the system non-linear. The state of the system
can no longer be predicted by linear theory and instead relies on non-linear dynamics.
When an instability has $\gamma = 0$ for any wave vector, we call it threshold\index{threshold} of the
associated instability. We also define $\gamma_{\max}$ for a given branch of dispersion
solutions as the maximum value of $\gamma$ for different wavenumbers ($\mathbf{k}$) and
for all propagation directions ($\theta$). $\omega_{max}$ and $\mathbf{k}_{\max}$ are
then defined as the values of $\omega$ and $\mathbf{k}$ corresponding to $\gamma =
\gamma_{\max}$ respectively. Mathematically this can be written as:
\begin{align}
\begin{split}
\gamma_{\max}^j = \mathrm{\max}\left(\gamma^j\left(\mathbf{k},\theta\right)\right)~~\forall \left(\mathbf{k}, \theta\right) \label{eq:gammamax}
\end{split}
\end{align}
where $j$ represents the branch of dispersion solutions for which $\gamma$ was computed.
In this thesis we consider ion temperature anisotropy as the only source of free energy.
For instability driven by ion temperature anisotropy, there are four distinct modes
depending on the anisotropy value ($R_{\rm p}$) and the direction of propagation
($\theta$). For parallel propagation ($\theta=0$), there are two modes: ion cyclotron
for $R_{\rm p} > 1$ and parallel firehose for $R_{\rm p} < 1$. In the oblique direction
($0 < \theta < 90$), the corresponding instabilities are mirror ($R_{\rm p} > 1$) and
oblique firehose ($R_{\rm p} < 1$). It is worth noting that neither of these two oblique
instabilities propagate. \Cref{tab:instab} gives a summary of all four instabilities
discussed.
\begin{table}[ht]
\centering
\caption[Microkinetic instabilities]{List of four temperature-anisotropy induced instabilities in plasma}
\begin{tabular}{ | p{0.25\linewidth} | p{0.25\linewidth} | p{0.25\linewidth}| }
\hline
Anisotropy Range & Parallel ($\omega_{\rm r} > 0$) & Oblique ($\omega_{\rm r} =
0$)\\
\hline
$R_{\rm p} > 1$ & Ion cyclotron & Mirror\\
\hline
$R_{\rm p} < 1$ & Parallel firehose & Oblique firehose\\
\hline
\end{tabular}
\label{tab:instab}
\end{table}
\subsection{Computing Growth Rates}\label{sec:cgr}
Computation of growth rates ($\gamma$) for any given value of the plasma parameters for
a given branch of instability means solving \Cref{eq:disp3} for all different values of
\textbf{k}. This is generally done by using a numerical dispersion solver. This project
used tables of $\gamma_{\max}$ as a function of the plasma parameters developed by
\citet{Maruca2012} using the software of \citet{Gary1993}. For our case, given a value
of $R_{\rm p}$ and $\beta_{\parallel \rm p}$, we check the table for values of ($R_{\rm
p}, \beta_{\parallel \rm p}$) that are closest to the input value and then use an
interpolation technique (cubic spline interpolation) to get the value at that exact
input point.
Once we get the $\gamma_{\max}$ value corresponding to an instability, we normalize it
to the cyclotron frequency of protons ($\gamma_{\max} \rightarrow
\gamma_{\max}/\Omega_{\rm cp}$). However, not all computed values of the growth rate
will have a significant effect on the plasma dynamics. As is expected if growth rate is
high, the instability grows faster and conversely it will take the instability a long
time to manifest or change the dynamics if it has a small growth rate. Thus, one can
think of $1/\gamma$ as a proxy for the amount of time it will take for the instability
to manifest itself. For a positive but an infinitesimal value of $\gamma_{\max}$, the
value of time will be much, much greater than the other characteristic time scales of
the system, which means the instability will never have time to significantly affect the
plasma. We thus define a cut-off\index{microinstability!cut-off} value of growth rates, in units of proton cyclotron
frequency, as:
\begin{align}
\gamma_{\max, \mathrm{cut\textrm{-}off}}^j & = 10^{-5}\,\Omega_\mathrm{{cp}} \label{eq:gammacutoff}
\end{align}
For any combination of ($R_{\rm p}, \beta_{\parallel \rm p}$) if the value of
$\gamma_{\max}/\Omega_\mathrm{{cp}}$ is less than $10^{-5}$ we essentially set those
values to zero and assume that the plasma is completely stable at those points. Choosing
this cut-off is a bit arbitrary. Different studies have chosen different threshold
values. For example \citet{Gary1999} and \citet{Gary2006} chose $\gamma =
10^{-2}\,\Omega_{\rm cp}$ as their cut-off, whereas both \citet{Hellinger2006} and
\citet{Klein2018} settled on cut-off value of $\gamma = 10^{-3}\,\Omega_{\rm cp}$. As
long as the cut-off value of $\gamma$ is significantly smaller than that of other
relevant time scales (see \Cref{sec:nlts,sec:data7}) it can comfortably be set to zero.
\section{Application of linear theory and observational evidence}\label{sec:app2}
The low density and extreme dynamics of space plasmas, such as solar wind and the
magnetosheath (see \Cref{sec:plas2}), ensure that they almost invariably deviate
substantially from local thermal equilibrium \citep{Marsch2006a,Verscharen2019}. For
example, even though the majority of solar wind ions are protons (ionized hydrogen) or
$\alpha$-particles (fully ionized helium), these two particle species rarely have equal
temperatures or bulk velocities \citep[see,
e.g.,][]{Feldman1974a,Marsch1982,Hefti1998,Kasper2008,Maruca2013a}. \Cref{fig:temp_1au}
shows the distribution of temperature for these two species using data from the Wind
spacecraft (\Cref{sec:wind} for more details on the dataset used). Furthermore, the VDF of
any given ion species often significantly departs from the entropically favored Maxwellian
functional form \citep{Feldman1973a,Feldman1974,Marsch1982b, Alterman2018}. Observations of
solar wind and magnetosheath from multiple spacecraft
\citep{Feldman1973b,Marsch1982b,Kasper2002} have shown that the protons exhibit distinct
kinetic temperatures, $T_{\rm \perp p}$ and $T_{\rm \parallel p}$. Both values of $R_{\rm p}
> 1$ and $R_{\rm p} < 1$ are commonly observed in the solar wind and in Earth's
magnetosheath, making it an ideal candidate for the application of linear Vlasov theory\index{Vlasov!linear Vlasov theory}.
\Cref{fig:aniso_wnd_mms} shows distribution of $R_{\rm p}$ for solar wind at 1\,au and
magnetosheath.
\begin{figure}
\begin{center}
\includegraphics[width=0.55\textwidth]{figures/chap2/proton_alpha_temp_dist_wnd.pdf}
\caption[Temperature distribution at 1\,au]{Distribution of proton and $\alpha$-temperatures at 1\,au. Vertical lines show the median temperature of each species.}
\label{fig:temp_1au}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.55\textwidth]{figures/chap2/proton_aniso_dis_mms_wnd.pdf}
\caption[$R_{\rm p}$ distribution at 1\,au and in magnetosheath]{Distribution of proton temperature anisotropy at 1\,au (red) and in the magnetosheath (blue). Vertical lines show the median values at each location.}
\label{fig:aniso_wnd_mms}
\end{center}
\end{figure}
As discussed in \Cref{sec:instab2} if $R_{\rm p}$ departs sufficiently from unity, it can
trigger a kinetic microinstability\index{microinstability}: a short-wavelength fluctuation with an exponentially
growing amplitude. The threshold $R_{\rm p}$-value for the onset of a proton
temperature-anisotropy instability depends on all plasma parameters (e.g., composition and
relative temperatures), depending most strongly on proton parallel beta ($\beta_{\parallel
\rm p}$). \Cref{fig:gamma_cntr} shows various thresholds on an ($R_{\rm p}, \beta_{\rm
\parallel p}$) plane for the four modes of instabilities. As is evident, at fixed $R_{\rm
p}$ a slight increment in $\beta_{\parallel \rm p}$ can lead to significant increase in the
growth rate.
\begin{figure}
\begin{center}
\includegraphics[width=1\textwidth]{figures/chap2/growth_rate_contours.pdf}
\caption[$\gamma$ Contours]{Contours of constant growth rates.}
\label{fig:gamma_cntr}
\end{center}
\end{figure}
These instabilities have threshold $R_{\rm p}$-values, which means that they can effectively
limit the degree to which proton temperature can depart from isotropy. If an unstable mode
grows and does not saturate, it eventually becomes nonlinear, continues to scatter particles
in phase space, and eventually drives the VDF toward local thermal equilibrium. Multiple
studies have analyzed large datasets from various spacecraft and under the assumptions of a
spatially homogeneous plasma and a bi-Maxwellian proton velocity distribution; such studies
have found that the joint distribution of $(\beta_{\parallel \rm p},R_{\rm p})$-values from
the interplanetary solar wind largely conform to the limits set by the instability
thresholds \citep{Gary2001,Kasper2002,Hellinger2006,Matteini2007}.
\Cref{fig:brazil_prob_wnd} shows the joint probability distribution of ($R_{\rm p},
\beta_{\parallel \rm p}$) and the thresholds\index{threshold} corresponding to different instability modes
for $\gamma_{\max}/\Omega_{\rm cp} = 10^{-2}$ \footnote{Please see \Cref{apdx:A} for more
details on how these figures were made and how thresholds were computed}. We can see that
the probability density decreases significantly as one moves closer to any of the threshold
values. A recent study by \citet{Maruca2018} confirmed the same effect in Earth's
magnetosheath, which is shown in \Cref{fig:brazil_prob_mms} (see \Cref{apdx:A} for examples
of a $(\beta_{\parallel \rm p},R_{\rm p})$-plot in other systems). Additional studies have
found that plasma with unstable $(\beta_{\parallel \rm p},R_{\rm p})$-values is
statistically more likely to exhibit enhancements in magnetic fluctuations \citep{Bale2009}
and proton temperature \citep{Maruca2011}. These findings suggest that the instabilities not
only regulate temperature anisotropy in space plasmas but, in doing so, play an integral
role in the large-scale evolution of the plasmas.
\begin{figure}
\begin{center}
\includegraphics[width=1\textwidth]{figures/chap2/brazil_prob_wnd.pdf}
\caption[Brazil-plot at 1\,au]{Plot of estimated probability density, $\Tilde{p}$ of
($R_{\rm p}, \beta_{\parallel \rm p}$) for solar wind at 1\,au (from \texttt{wnd}
dataset, see \Cref{chap:chap4}) and thresholds associated with different
instabilities for threshold value of $\gamma_{\max}/\Omega_{\rm cp} = 10^{-2}$.}
\label{fig:brazil_prob_wnd}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=1\textwidth]{figures/chap2/brazil_prob_mms.pdf}
\caption[Brazil-plot in magnetosheath]{Plot of estimated probability density,
$\Tilde{p}$ of ($R_{\rm p}, \beta_{\parallel \rm p}$) for Earth's magnetosheath
(from \texttt{mms} dataset, see \Cref{chap:chap4}) and thresholds associated with
different instabilities for threshold value of $\gamma_{\max}/\Omega_{\rm cp} =
10^{-2}$.}
\label{fig:brazil_prob_mms}
\end{center}
\end{figure}
The empirical studies of $(\beta_{\parallel \rm p},R_{\rm p})$-distributions --- especially
that by \citet{Matteini2007} --- indicate that the instabilities globally limit proton
temperature anisotropy and affect the large-scale thermodynamics of expanding solar wind
plasma. Nevertheless, the instabilities themselves act on far smaller scales. Indeed,
\cite{Osman2012} found that unstable $(\beta_{\parallel \rm p},R_{\rm p})$-values are
statistically more likely to exhibit enhanced values of the partial variance of increments
(PVI)\index{PVI}, which is an indicator of intermittent structure (see \Cref{sec:intmt}). This result
suggests that long-wavelength turbulence may play a substantial role in generating the local
plasma conditions that drive these microinstabilities. Also, advancements made in numerical
simulation by \citet{Servidio2012a, Greco2012, Servidio2015}, with corroboration from space
plasma observations \citep{Marsch1992, Sorriso-Valvo1999, Osman2011, Osman2012, Kiyani2009}
show the importance of intermittency in interpretation of these observations. We discuss
these in more detail in \Cref{chap:chap3}.
\section{Limitations of linear theory}\label{sec:conc2}
Though linear theory works well for plasma with a homogeneous background, when it comes to
its application to study the characteristics of space plasmas, the method is not without
caveats. Multiple studies have shown space plasma to be highly structured and thus
inhomogeneous \citep{Burlaga1968, Tsurutani1979, Ness2001, Osman2012, Osman2012a,
Greco2012}. In fact, by all accounts, inhomogeneity is ubiquitously present in space plasma,
and thus any study of instabilities in plasma should take into account the inhomogeneity of
the background among variation in other parameters.
Consequently, use of linear theory for such studies of course presents a theoretical
inconsistency in the application of computed instability thresholds\index{threshold} to study the properties
of plasma because of the underlying disparity between the assumptions of linear theory and
the observed space plasma. However, several studies over the last three decades have
presented empirical evidence of agreement between the observations and theoretical
predictions \citep{Gary1991, Gary1994, Gary2001, Gary2006, Kasper2002, Hellinger2006,
Maruca2011, Maruca2012, Maruca2018}. These studies strongly suggest that linear instability
thresholds are indeed efficient in restricting the plasma/plasma VDF in a narrow region of
($\beta_{\parallel \rm p}, R_{\rm p}$)-plane, inhibiting the excursion of plasma VDFs to
extreme anisotropy regions at high $\beta_{\parallel \rm p}$. Although limitations on
spatial and temporal resolution using present-day spacecraft make it difficult to directly
demonstrate the existence of such instabilities in space plasmas, work done by,
\citet{Bale2009, He2011, Podesta2013, Jian2009, Jian2010, Jian2014, Klein2014, Telloni2016,
Gary2016} among others provide indirect evidence for the presence of various different
instabilities. More details can be found in \citet{Verscharen2019} and references therein.
Given these limitations of linear theory and its application, we have to look into the
non-linear processes and study how those processes affect the dynamics of the plasma.
\Cref{chap:chap3} introduces some the non-linear processes in plasmas and discusses how they
affect the dynamics.