physics0501066/fp.tex
1: % Copy for PPPL report and Computer Physics Reports (Nov 12, 1985)
2: % Revised Jan 13, 1986
3: % Revised again Jan 27, 1986
4: \documentstyle[aipmod,eqalign,graphicx,times]{article}
5: %\mag=\magstep1 \advance\topmargin -1truein
6: %\advance\oddsidemargin -.65truein \evensidemargin\oddsidemargin
7: %\def\baselinestretch{1.2}
8: \advance \textwidth by 1in
9: \advance \oddsidemargin by -0.5in
10: \advance \textheight by 1in
11: \advance \topmargin by -0.5in
12: \let\sc=\small
13: \font\tenmib=cmmib10 scaled \magstep1 \font\tensfb=cmssbx10 scaled \magstep1 
14: \def\dv{d^3{\bf v}}\def\da{d^2\!{\bf A}}\def\dvp{d^3{\bf p}}
15: \def\bphi{\hbox{\tenmib \char'036 }}
16: \def\bsigma{\hbox{\tenmib \char'033 }}
17: \def\abs#1{\left|#1\right|}\def\ave#1{\left<#1\right>}
18: \def\sigin{\Sigma_{\rm in}}\def\sigout{\Sigma_{\rm out}}
19: \def\sign{\mathop{\rm sign}\nolimits}
20: \def\bGamma{\hbox{\tenbf \char'000 }}\def\mat#1{\hbox{\tensfb #1}}
21: \def\secref#1{sec.~\ref{#1}}\def\Secref#1{Section~\ref{#1}}
22: \def\secsref#1{secs.~\ref{#1}}\def\Secsref#1{Sections~\ref{#1}}
23: \def\appref#1{appendix~\ref{#1}}\def\Appref#1{Appendix~\ref{#1}}
24: \def\appsref#1{appendices~\ref{#1}}\def\Appsref#1{Appendices~\ref{#1}}
25: \def\eqref#1{eq.~(\ref{#1})}\def\Eqref#1{Equation~(\ref{#1})}
26: \def\eqsref#1{eqs.~(\ref{#1})}\def\Eqsref#1{Equations~(\ref{#1})}
27: \def\figref#1{fig.~\ref{#1}}\def\Figref#1{Figure~\ref{#1}}
28: \def\tabref#1{table~\ref{#1}}\def\Tabref#1{Table~\ref{#1}}
29: \def\Eqlab#1{\eqn(\eqlab{#1})}
30: \def\Refref#1{Reference~\citenum{#1}}\def\refref#1{ref.~\citenum{#1}}
31: \def\grapprox{\mathrel{\rlap{\lower3.5pt\hbox{$\mathchar"218$}}\raise 1pt
32: 				\hbox{$\mathchar"13E$}}}
33: \def\lsapprox{\mathrel{\rlap{\lower3.5pt\hbox{$\mathchar"218$}}\raise 1pt
34: 				\hbox{$\mathchar"13C$}}}
35: \def\grls{\mathrel{\rlap{\lower2.5pt\hbox{$\mathchar"13C$}}\raise2.5pt
36: 						\hbox{$\mathchar"13E$}}}
37: \def\lsgr{\mathrel{\rlap{\lower2.5pt\hbox{$\mathchar"13E$}}\raise2.5pt
38: 						\hbox{$\mathchar"13C$}}}
39: \def\fract#1/#2{{\textstyle{#1\over #2}}}
40: \def\half{\fract1/2}
41: \hyphenation{max-well-ian}
42: 
43: \begin{document}
44: \date{PPPL--2290 (Nov.~1985)\\
45: Comp.\ Phys.\ Rep.\ {\bf 4}(3--4), 183--244 (Aug.~1986)}
46: \title{Fokker--Planck and Quasilinear Codes\thanks
47: { Presented at the 3rd European Workshop on Problems in the Numerical Modeling
48:   of Plasmas,  Varenna, Italy, September 10--13, 1985.
49:   Reprinted in {\em Problems in the Numerical Modeling of Plasmas},
50:   edited by K. Appert (1986).
51: }} 
52: \author{Charles F. F. Karney\\
53: Plasma Physics Laboratory, Princeton University\\
54: P.O.\ Box 451\\
55: Princeton, New Jersey 08544--0451, U.S.A.}
56: \maketitle
57: 
58: \begin{abstract}
59: The interaction of radio-frequency waves with a plasma is described by
60: a Fokker--Planck equation with an added quasilinear term.  Methods for
61: solving this equation on a computer are discussed.
62: \end{abstract}
63: %\tableofcontents
64: \section{Introduction}\label{intro}
65: 
66: In this paper, I will concentrate on those Fokker--Planck models which
67: are most useful for the study of rf-driven currents \cite{Fisch1}.  I
68: will therefore take the plasma to be azimuthally symmetric about the
69: magnetic field and homogeneous (representative of the central portion
70: of a tokamak plasma).  The Fokker--Planck equation then reduces to an
71: equation in time and two velocity (or momentum) dimensions only.  This
72: simplified model yields a wealth of interesting physics and furthermore
73: illustrates the main numerical problems encountered in more complicated
74: situations.  In addition to the collision term, the equation will
75: include the effects of externally injected rf power via a quasilinear
76: diffusion term, and a dc electric field.  (The electric field arises
77: whenever the current is time-varying, e.g., during current ramp-up.)
78: Because the wave interacts with very fast electrons, relativistic
79: effects are also considered.  In addition, the adjoint method for
80: solving for moments of the Fokker--Planck equation is discussed.  This
81: method allows for a great reduction (by orders of magnitude) in the
82: amount of computer time required.
83: 
84: The paper is divided into three parts:
85: In the first part of the paper, I give the formulation of the
86: Fokker--Planck equation.  In \secref{prelim}, the
87: Fokker--Planck equation and the coordinate systems are introduced.
88: The collision operator and approximations to it are given in
89: \secsref{coll} and \ref{coll-approx}.  Corresponding expressions for 
90: the  quasilinear diffusion
91: operator are given in \secref{ql}.  The next part of the paper
92: describes the numerical solution of the equation.  Its boundary
93: conditions are considered in \secref{bcs}.  \Secsref{spatial}
94: and \ref{time} describe the spatial and temporal differencing  of the
95: equation.  In \secref{steady-state}, we describe techniques for
96: obtaining the time asymptotic solution to the equation.  The last part
97: of the paper describes the incorporation of relativistic effects
98: (\secref{relativity}) and the adjoint method for solving the
99: Fokker--Planck equation (\secref{adjoint}).
100: 
101: The numerical methods presented here are those used in the
102: Fokker--Planck code used by the author.  A word about the lineage of
103: this code is in order:  Fokker--Planck codes were developed at Livermore
104: by Killeen {\sl et al.}\ \cite{Killeen1,Killeen2} for the study of
105: mirror-machine plasmas.  The latest stage in the development of these
106: codes is {\sc FPPAC} \cite{McCoy} which is a two-dimensional
107: multispecies nonlinear Fokker--Planck package.  The Livermore code was
108: extensively modified by Winsor and Fallon for the study of runaway
109: electrons, which was undertaken by Kulsrud {\sl et al.}\ \cite{Kulsrud}.
110: This code has been used by the author in various studies of current
111: drive beginning with lower hybrid current drive \cite{Karney-lh}.  Over
112: the years several further modifications have been made, although the
113: basic structure of the code is the same as that of Kulsrud {\sl et al.}
114: 
115: The assumption of a homogeneous magnetic field is warranted in the study
116: of rf heating and current drive in tokamaks if the rf interacts only
117: with circulating particles.  This is often not the case (for example
118: during ion- and electron-cyclotron heating) in which case proper
119: account should be taken of trapped particles.  This has been done in
120: so-called {\sl bounce-averaged} codes \cite{Cutler,Kerbel,Matsuda} in
121: which the distribution function is averaged over the bounce motion of
122: the trapped particles.  In tokamaks this leads to a modification of the
123: coefficients appearing in the Fokker--Planck equation but the numerical
124: treatment of the equation is largely unaltered.  In machines with more
125: complicated particle orbits, the distribution may be a multivalued
126: function of the velocity coordinates.  This occurs in tandem mirrors
127: where there is more than one population of trapped particles.  In this
128: case, special techniques are required \cite{Matsuda}.
129: 
130: \section{Preliminaries}\label{prelim}
131: 
132: \subsection{The Fokker--Planck equation}
133: 
134: We write the Fokker--Planck equation for the electrons $e$ as
135: 	$${\partial f_e\over\partial t}-\sum_s C(f_e,f_s) +
136: \nabla\cdot{\bf S}_w+{q_e{\bf E}\over m_e}\cdot\nabla f_e = 0, \Eqlab{fp-1}$$
137: 	where $q_s$ and $m_s$ are the charge and mass of species $s$,
138: $C(f_a,f_b)$ is the collision term for species $a$ colliding off
139: species $b$, the sum extends over all the species of the plasma
140: (typically electrons and ions), ${\bf S}_w$ is the wave ($w$)-induced
141: quasilinear flux, and ${\bf E} =E{\bf \hat v}_\parallel$ is the
142: electric field (assumed to be parallel to the magnetic field).  The
143: quantity $q_s$ carries the sign of the charge, thus $q_e=-e$.  The
144: subscripts $\parallel$ and $\perp$ refer to the directions parallel and
145: perpendicular to the magnetic field.  The
146: $\nabla\equiv\partial/\partial{\bf v}$ operator operates in velocity
147: space.
148: 
149: Because collisions in a plasma are primarily due to small-angle
150: scattering, the collision term can be written as the divergence of a
151: flux
152: 	$$C(f_a,f_b)=-\nabla\cdot {\bf S}_c^{a/b},$$
153: in which case \eqref{fp-1} can be expressed as
154: 	$${\partial f_e\over\partial t}+
155: \nabla\cdot{\bf S}= 0, \Eqlab{fp-2}$$
156: where 
157: 	$${\bf S}={\bf S}_c+{\bf S}_w+
158: {\bf S}_e$$
159: 	is the total flux in velocity space, and
160: $$\eqalignno{{\bf S}_c&=\sum_s{\bf S}_c^{e/s},&(\eqlab{s-coll})\cr
161: {\bf S}_e&={q_e{\bf E}\over m_e}f_e,&(\eqlab{s-e})\cr}$$
162: are the collisional ($c$)- and electric-field ($e$)-induced electron fluxes.
163: 
164: From \eqref{fp-2} we can derive the conservation laws
165: 	$$\eqalignno{
166: {\partial\over\partial t}
167: \int_V f_e\,\dv+\int_A {\bf S}\cdot \da&=0,&(\eqlab{cons}a)\cr
168: {\partial\over\partial t}
169: \int_V m_e{\bf v}f_e\,\dv+
170: \int_A m_e{\bf v}{\bf S}\cdot \da&=\int_V m_e{\bf S}\,\dv,
171: &(\ref{cons}b)\cr
172: {\partial\over\partial t}
173: \int_V {m_ev^2\over2} f_e\,\dv+
174: \int_A {m_ev^2\over2}{\bf S}\cdot \da&=\int_V m_e{\bf v\cdot
175: S}\,\dv,&(\ref{cons}c)\cr}$$
176: 	where $V$ is some volume in velocity space and $A$ is its
177: boundary.  These equations are statements of conservation of number,
178: momentum, and energy.
179: 
180: Typically, two types of terms appear in ${\bf S}$:  a diffusion term and
181: a friction term
182: 	$${\bf S}=-\mat D\cdot \nabla f_e+ {\bf F} f_e.\Eqlab{diff-frict}$$
183: 	The wave term is purely diffusive so that ${\bf F}_w=0$, while
184: the electric field term is nondiffusive: $\mat D_e=0$,
185: 	$${\bf F}_e={q_e{\bf E}\over m_e}.\Eqlab{e-flux}$$
186: 
187: \subsection{Coordinate systems}
188: 
189: Because of azimuthal symmetry, $f_e$ is independent of $\phi$ the angle
190: about the magnetic field.  Two coordinate systems suggest themselves:
191: the cylindrical coordinate system $(v_\perp,v_\parallel,\phi)$ and the
192: spherical coordinate system $(v,\theta,\phi)$; see \figref{coord-fig}.
193: These are related by
194: 	$$\eqalign{v^2&=v_\perp^2+v_\parallel^2,\cr
195: \cos\theta&=v_\parallel/v.\cr}$$
196: 	Both of these coordinate systems are useful.  In cylindrical
197: coordinates (assuming azimuthal symmetry) \eqref{diff-frict} gives
198: 	$$\eqalignno{\nabla\cdot{\bf S}&=
199: {1\over v_\perp}{\partial\over\partial v_\perp}v_\perp S_\perp
200: +{\partial\over\partial v_\parallel}S_\parallel,&(\eqlab{cylindrical}a)\cr
201: S_\perp&=
202: -D_{\perp\perp} {\partial f_e\over\partial v_\perp}
203: -D_{\perp\parallel}{\partial f_e\over\partial v_\parallel}
204: +F_{\perp}f_e,&(\ref{cylindrical}b)\cr
205: S_\parallel&=
206: -D_{\parallel\perp} {\partial f_e\over\partial v_\perp}
207: -D_{\parallel\parallel}{\partial f_e\over\partial v_\parallel}
208: +F_{\parallel}f_e.&(\ref{cylindrical}c)\cr}$$
209: 	Similarly, in spherical coordinates we have
210: 	$$\eqalignno{\nabla\cdot{\bf S}&=
211: {1\over v^2}{\partial\over\partial v}v^2S_v+
212: {1\over v\sin\theta}{\partial\over\partial \theta}\sin\theta S_\theta,
213: &(\eqlab{spherical}a)\cr
214: S_v&=
215: -D_{v v} {\partial f_e\over\partial v}
216: -D_{v \theta}{1\over v}{\partial f_e\over\partial \theta}
217: +F_{v}f_e,&(\ref{spherical}b)\cr
218: S_\theta&=
219: -D_{\theta v} {\partial f_e\over\partial v}
220: -D_{\theta \theta}{1\over v}{\partial f_e\over\partial \theta}
221: +F_{\theta}f_e.&(\ref{spherical}c)\cr}$$
222: 	Transformations between $\mat D$ and $\bf F$ expressed in the
223: two coordinate systems may be achieved by
224: 	$$\pmatrix{D_{\perp\perp}\cr D_{\perp\parallel}\cr
225: D_{\parallel\perp}\cr D_{\parallel\parallel}\cr}
226: =\mat M\cdot \pmatrix{D_{v v}\cr D_{v \theta}\cr
227: D_{\theta v}\cr D_{\theta \theta}\cr}
228: \eqn(\eqlab{transform}a)$$
229: 	and
230: 	$$\pmatrix{F_{\perp}\cr F_{\parallel}\cr}=
231: \mat N\cdot \pmatrix{F_{v}\cr F_{\theta}\cr},
232: \eqn(\ref{transform}b)$$
233: 	where
234: 	$$\eqalign{\mat M=\mat M^{-1}&=
235: \pmatrix{s^2&sc&sc&c^2\cr   sc&-s^2&c^2&-sc\cr
236:          sc&c^2&-s^2&-sc\cr c^2&-sc&-sc&s^2\cr}\cr
237: \mat N=\mat N^{-1}&=
238: \pmatrix{s&c\cr c&-s\cr},\cr}$$
239: 	and here we have abbreviated $s=\sin\theta$ and $c=\cos\theta$.
240: The collision term is most conveniently expressed in spherical
241: coordinates and \eqsref{transform} allow us to transform this term to
242: cylindrical coordinates.  On the other hand, the rf and electric field
243: terms are written most naturally in cylindrical coordinates, and this
244: equation also enables us to express these terms in spherical
245: coordinates.
246: 
247: In the case of the collision operator, $\mat D$ and $\bf F$ are given
248: in terms of the gradients of potentials
249: $$\mat D \propto \nabla\nabla\psi,\qquad {\bf F}\propto\nabla\phi.$$
250: 	In cylindrical coordinates (with azimuthal symmetry), the
251: relevant components of $\mat D$ and $\bf F$ are easy to calculate---we
252: just take the corresponding derivatives of $\psi$ and $\phi$.  In
253: spherical coordinates we have
254: 	$$\eqalignno{(\nabla\nabla\psi)_{vv}&=
255: {\partial^2 \psi\over\partial v^2},
256: &(\eqlab{spherical-comp}a)\cr
257: (\nabla\nabla\psi)_{v\theta}=(\nabla\nabla\psi)_{\theta v}&=
258: {1\over v}{\partial^2 \psi\over\partial v\partial \theta}
259: -{1\over v^2}{\partial \psi\over\partial \theta},&(\ref{spherical-comp}b)\cr
260: (\nabla\nabla\psi)_{\theta\theta}&={1\over v}{\partial \psi\over\partial v}
261: +{1\over v^2}{\partial^2 \psi\over\partial \theta^2},&(\ref{spherical-comp}c)\cr
262: (\nabla\phi)_v&={\partial \phi\over\partial v},&(\ref{spherical-comp}d)\cr
263: (\nabla\phi)_\theta&={1\over v}{\partial \phi\over\partial \theta}.
264: &(\ref{spherical-comp}e)\cr}$$
265: 
266: Several important quantities are given in terms of velocity-space
267: moments of the distribution function.  Three-dimensional velocity
268: space integrations can be carried out in cylindrical coordinates using
269: 	$$\int f({\bf v})\,\dv
270: =\int_0^\infty\! dv_\perp\int_{-\infty}^\infty\! dv_\parallel
271: \, 2\pi v_\perp f(v_\perp,v_\parallel)
272: \Eqlab{int-cyl}$$
273: 	and in spherical coordinates using
274: 	$$\int f({\bf v})\,\dv
275: =\int_0^\infty\! dv\int_0^\pi\! d\theta\, 2\pi v^2 f(v,\theta)\sin\theta.
276: \Eqlab{int-spher}$$
277: 
278: \subsection{Legendre harmonics}
279: 
280: 	It is sometimes useful to decompose the distribution function
281: and the potentials into Legendre harmonics $P_l(\cos\theta)$.  We write
282: this as
283: 	$$f(v,\theta)=\sum_{l=0}^\infty f^{(l)}(v)P_l(\cos\theta),
284: \Eqlab{legend-comp}$$
285: where
286: $$f^{(l)}(v)={2l+1\over 2}
287: \int_0^\pi f(v,\theta) P_l(\cos\theta)\sin\theta\,d\theta.
288: \Eqlab{legend-decomp}$$
289: 	The Legendre polynomials may be evaluated on a computer using
290: the recurrence relation \cite{Abramowitz}
291: 	$$P_0(\mu)=1,\qquad
292: 	(l+1)P_{l+1}(\mu)=(2l+1)\mu P_l(\mu)-lP_{l-1}(\mu).$$
293: 
294: \subsection{Definitions}
295: 
296: Finally, we define some of the other quantities that we encounter in
297: this paper.  The thermal velocity of species $s$ is given by
298: $$v_{ts}=\sqrt{T_s\over m_s},\Eqlab{vts-max}$$
299: 	where $T_s$ is the temperature of species $s$.  The thermal
300: collision frequency for the electrons is
301: 	$$\nu_{te}=\tau_{te}^{-1}={\Gamma^{e/e}\over 
302: v_{te}^3},\Eqlab{nute}$$
303: where
304: $$\Gamma^{a/b}={n_b q_a^2 q_b^2 \ln \Lambda^{a/b}\over 4\pi
305: \epsilon_0^2 m_a^2},$$
306: 	and $n_s$ is the number density of species $s$, $\epsilon_0$ is
307: the dielectric constant of free space, and $\ln \Lambda^{a/b}$ is the
308: Coulomb logarithm.  The distributions are normalized so that
309: 	$$\int f_s({\bf v})\,\dv=n_s.$$
310: 	In particular, the Maxwellian distribution is
311: $$\eqalignno{f_{sm}(v)&=n_s\biggl({m_s\over 2\pi T_s}\biggr)^{3/2}
312:  \exp(-\half m_sv^2/T_s),\cr
313: &=n_s{1\over(2\pi v_{ts}^2)^{3/2}}
314:  \exp(-\half v^2/v_{ts}^2).&
315: (\eqlab{max})\cr}$$
316: 
317: In discussing the applications to rf current drive, there are two
318: quantities in which we will be interested: the electron current density
319: 	$$J=\int q_ev_\parallel f_e({\bf v})\,\dv\Eqlab{curr-def}$$
320: and the rf power absorbed per unit volume by the plasma
321: 	$$P=\int m_e{\bf v}\cdot{\bf S}_w\,\dv.\Eqlab{pd-def}$$
322: 	The efficiency of rf current drive is usually given as the
323: ratio $J/P$.
324: 
325: We shall use {\sc S.I.} units throughout this paper except that we will
326: measure temperature in units of energy.
327: 
328: \section{Collision Operator}\label{coll}
329: \subsection{The Landau collision operator}
330: 
331: The collision flux is given by the Landau collision integral
332: \cite{Landau}
333: 	$${\bf S}_c^{a/b}={q_a^2 q_b^2 \over 8\pi\epsilon_0^2 m_a}
334: \ln \Lambda^{a/b}\int \mat U({\bf u})\cdot
335: \biggl(
336: {f_a({\bf v})\over m_b} {\partial f_b({\bf v}')\over\partial {\bf v}'}-
337: {f_b({\bf v}')\over m_a} {\partial f_a({\bf v})\over\partial {\bf v}}
338: \biggr) \,\dv',\Eqlab{coll-land}$$
339: 	where
340: 	$$
341: \mat U({\bf u})={u^2\mat I - {\bf uu}\over u^3},\qquad
342: {\bf u}={\bf v}-{\bf v}'.$$
343: 	The formula for the Coulomb logarithm $\ln \Lambda^{a/b}$
344: is given in text books and the {\sl NRL Plasma Formulary} \cite{NRL}.  Because it
345: is so insensitive to plasma parameters, in many cases it is adequate to
346: take it to be a constant equal to $15$.  In any case, it is required
347: that $\ln\Lambda^{a/b}=\ln\Lambda^{b/a}$.  The Landau collision operator
348: conserves number, momentum, and energy, i.e.,
349: 	$$\eqalignno{\int C(f_a,f_b)\,\dv&=0,
350: &(\eqlab{land-cons}a)\cr
351: \int m_a{\bf v}C(f_a,f_b)\,\dv+
352: \int m_b{\bf v}C(f_b,f_a)\,\dv&=0,&(\ref{land-cons}b)\cr
353: \int {m_a v^2\over2}C(f_a,f_b)\,\dv+
354: \int {m_b v^2\over2}C(f_b,f_a)\,\dv&=0.&(\ref{land-cons}c)\cr}$$
355: 
356: There is an error on the order of $1/\ln\Lambda^{a/b}$ in the Landau
357: collision operator.  However, because it has so many ``nice''
358: properties---the conservation laws of \eqsref{land-cons}, an
359: $H$-theorem, etc.---it is customary to regard \eqref{coll-land} as
360: being exact.
361: 
362: \subsection{Rosenbluth potentials}
363: 
364: \Eqref{coll-land} is the most useful form for the collision operator
365: for analytical work.  However, it is not in a convenient form for
366: numerical computations.  Suppose we represent the distribution functions
367: on an $N\times N$ grid (assuming azimuthal symmetry).  Then evaluation of
368: \eqref{coll-land} entails $O(N^4)$ computations, because it entails a
369: two-dimensional integral (over ${\bf v}'$) which must be carried out at
370: each grid location.  Fortunately, substantial savings may be realized by
371: using an equivalent representation in terms of Rosenbluth potentials
372: \cite{Rosenbluth,Trubnikov1}.  Here we use the slightly more convenient
373: notation of Trubnikov \cite{Trubnikov}.  We define two potentials
374: 	$$\eqalignno{
375: \phi_b({\bf v})&=-{1\over 4\pi}
376: \int {f_b({\bf v}')\over\abs{{\bf v}-{\bf v}'}}\,\dv',
377: &(\eqlab{pot}a)\cr
378: \psi_b({\bf v})&=-{1\over 8\pi}
379: \int \abs{{\bf v}-{\bf v}'}f_b({\bf v}')\,\dv'.
380: &(\ref{pot}b)\cr}$$
381: 	These are called potentials because they
382: satisfy Poisson's equations in velocity space
383: $$\nabla^2\phi_b({\bf v})=f_b({\bf v}),\qquad
384: \nabla^2\psi_b({\bf v})=\phi_b({\bf v}).$$
385: 	In terms of these potentials, \eqref{coll-land} becomes
386: 	$$\eqalignno{{\bf S}_c^{a/b}&=
387: -\mat D_c^{a/b}\nabla f_a({\bf v})+{\bf F}_c^{a/b}f_a({\bf v}),&
388: (\eqlab{coll-ros}a)\cr
389: \mat D_c^{a/b}&=-{4\pi\Gamma^{a/b}\over n_b}
390: \nabla\nabla\psi_b({\bf v}),&(\ref{coll-ros}b)\cr
391: {\bf F}_c^{a/b}&=-{4\pi\Gamma^{a/b}\over n_b}
392: {m_a\over m_b}\nabla\phi_b({\bf v}).&(\ref{coll-ros}c)\cr}$$
393: 	(An equivalent form of this equation is given by Rosenbluth
394: {\sl et al.}\ \cite{Rosenbluth} which contains a term of the form
395: 	$$\nabla\cdot[f_a({\bf v})\nabla\nabla\psi_b({\bf v})].$$
396: 	This form is used in several numerical codes
397: \cite{Killeen1,Killeen2,McCoy}, even though more derivatives of
398: $\psi_b$ must be taken.  One form may be derived from the other by
399: noting that $\nabla^2\psi_b=\phi_b$.)
400: 
401: 	There is an efficient method for calculating the Rosenbluth
402: potentials.  This involves decomposing $f_b$ in Legendre harmonics
403: \eqref{legend-decomp}.  Then we have \cite{Rosenbluth}
404: 	$$\eqalignno{\phi_b^{(l)}(v)&=-{1\over 2l+1}
405: \biggl[\int_0^v {v'{}^{l+2}\over v^{l+1}} f_b^{(l)}(v')\,dv'
406: +\int_v^\infty {v^l\over v'{}^{l-1}}f_b^{(l)}(v')\,dv'\biggr],&
407: (\eqlab{pot1}a)\cr
408: \psi_b^{(l)}(v)&={1\over 2(4l^2-1)}
409: \biggl[\int_0^v {v'{}^{l+2}\over v^{l-1}}
410: \biggl(1-{l-\half\over l+\fract3/2}{v'{}^2\over v^2}\biggr)
411: f_b^{(l)}(v')\,dv'\cr&\qquad\qquad\qquad
412: +\int_v^\infty {v^l\over v'{}^{l-3}}
413: \biggl(1-{l-\half\over l+\fract3/2}{v^2\over v'{}^2}\biggr)
414: f_b^{(l)}(v')\,dv'\biggr].&
415: (\ref{pot1}b)\cr}$$
416: 	Let us assume that $f_b$ may be represented by $K$ Legendre
417: harmonics (i.e., the upper limit in the sum in \eqref{legend-comp} is
418: $K-1$).  Then the calculation of $f_b^{(l)}(v)$ from $f_b({\bf v})$
419: using \eqref{legend-decomp} takes $O(N^2)$ computations for each $l$ or
420: $O(KN^2)$ computations altogether.  Given $f_b^{(l)}(v)$, the
421: calculation of $\phi_b^{(l)}(v)$ and $\psi_b^{(l)}(v)$ using
422: \eqsref{pot1} takes $O(N)$ computations for each $l$.  The
423: calculation of $\phi({\bf v})$ and $\psi({\bf v})$ takes a further
424: $O(KN^2)$ step.  Overall the number of steps is therefore $O(KN^2)$.
425: Often, we can take $K$ to be quite small (usually $K<10$), and in any
426: case we have $K\le N$, so we can compute the collision term much more
427: economically than using the Landau operator directly.
428: 
429: \section{Approximations to the Collision Operator}\label{coll-approx}
430: \subsection{Isotropic background}
431: 
432: If the background distribution is isotropic $f_b({\bf v})=f_b(v)$, then
433: so too are $\phi$ and $\psi$.  The collision term is then from
434: \eqsref{coll-ros}, (\ref{pot1}), and (\ref{spherical-comp})
435: 	$$\eqalignno{S_{cv}^{a/b}&=
436: -D_{cv v}^{a/b} {\partial f_a\over\partial v}
437: +F_{cv}^{a/b}f_a,&(\eqlab{isotropic}a)\cr
438: S_{c\theta}^{a/b}&=
439: -D_{c\theta \theta}^{a/b}{1\over v}{\partial f_a\over\partial \theta}
440: ,&(\ref{isotropic}b)\cr}$$
441: 	where
442: 	$$\eqalignno{D_{cvv}^{a/b}&=
443: {4\pi\Gamma^{a/b}\over 3n_b}
444: \biggl(\int_0^v {v'{}^4\over v^3}f_b(v')\,dv'
445: +\int_v^\infty v' f_b(v')\,dv'\biggr),&(\eqlab{iso-gen}a)\cr
446: D_{c\theta\theta}^{a/b}&=
447: {4\pi\Gamma^{a/b}\over 3n_b}
448: \biggl(\int_0^v {v'{}^2\over 2v^3}(3v^2-v'{}^2)f_b(v')\,dv'
449: +\int_v^\infty v' f_b(v')\,dv'\biggr),&(\ref{iso-gen}b)\cr
450: F_{cv}^{a/b}&=
451: -{4\pi\Gamma^{a/b}\over 3n_b}
452: {m_a\over m_b}
453: \int_0^v{3v'{}^2\over v^2}f_b(v')\,dv'.&(\ref{iso-gen}c)\cr}$$
454: 
455: \subsection{The high-velocity limit}
456: 
457: If $v$ is much greater than the thermal velocity of particles of
458: species $b$, the indefinite limits in \eqsref{iso-gen} may be replaced
459: by infinity to give
460: 	$$\eqalignno{D_{cvv}^{a/b}&=
461: \Gamma^{a/b} {v_{tb}^2\over v^3}
462: ,&(\eqlab{iso-high}a)\cr
463: D_{c\theta\theta}^{a/b}&=
464: \Gamma^{a/b} {1\over 2v}\biggl(1-{v_{tb}^2\over v^2}\biggr)
465: ,&(\ref{iso-high}b)\cr
466: F_{cv}^{a/b}&=
467: -\Gamma^{a/b} {m_a\over m_b}
468: {1\over v^2},&(\ref{iso-high}c)\cr}$$
469: 	where the thermal velocity is defined for an arbitrary isotropic
470: distribution as
471: 	$$v_{ts}^2={4\pi\over 3n_s}\int_0^\infty v^4 f_s(v)\,dv.
472: \Eqlab{vts}$$
473: [For a Maxwellian distribution, \eqref{max}, this reduces to the usual
474: expression \eqref{vts-max}.]
475: 
476: \subsection{Maxwellian background}
477: 
478: If the background distribution is Maxwellian \eqref{max}, the integrals
479: in \eqref{iso-gen} can be carried out to give \cite{Trubnikov}
480: \def\erf{\mathop{\rm erf}\nolimits}%
481: 	$$\eqalignno{D_{cvv}^{a/b}&={\nu_\parallel^{a/b}\over 2}v^2
482: ={\Gamma^{a/b}\over 2v}\biggl({\erf(u)\over u^2}-{\erf'(u)\over u}
483: \biggr),&(\eqlab{iso-max}a)\cr
484: D_{c\theta\theta}^{a/b}&={\nu_\perp^{a/b}\over 4}v^2
485: ={\Gamma^{a/b}\over 4v}\biggl(\biggl(2-{1\over u^2}\biggr)\erf(u)
486: +{\erf'(u)\over u}\biggr),&(\ref{iso-max}b)\cr
487: F_{cv}^{a/b}&=-{m_a\over m_a+m_b}\nu_s^{a/b}v
488: =-{\Gamma^{a/b}\over v^2}{m_a\over m_b}\bigl[\erf(u)-u\erf'(u)
489: \bigr],&(\ref{iso-max}c)\cr}$$
490: 	where
491: $$\eqalign{\erf(u)&={2\over\sqrt\pi}\int_0^u\exp(-x^2)\,dx,\cr
492: \erf'(u)&={2\over\sqrt\pi}\exp(-u^2),\cr
493: u&={v\over\sqrt2 v_{tb}}.\cr}$$
494: 	The parallel diffusion rate $\nu_\parallel^{a/b}$,
495: perpendicular diffusion rate $\nu_\perp^{a/b}$, and the slowing down
496: diffusion rate $\nu_s^{a/b}$ are the same as those defined in the {\sl
497: NRL Plasma Formulary} \cite{NRL}.  [However, the {\sl NRL Plasma
498: Formulary} (1983 edition) has an incorrect formula for the collision
499: operator with a Maxwellian background.]
500: 
501: 	For $u>0$, $\erf(u)$ is given approximately by \cite{Abramowitz}
502: 	$$\erf(u)=1-\exp(-u^2)\sum_{k=1}^5 a_k t^k,$$
503: 	where
504: 	$$t=1/(1+pu)$$ and
505: 	$$\eqalign{p&=0.32759\,11,\cr
506: 	a_1&=0.25482\,9592,\qquad a_2=-0.28449\,6736,\cr
507: 	a_3&=1.42141\,3741,\qquad a_4=-1.45315\,2027,\cr
508: 	a_5&=1.06140\,5429.\cr}$$
509: 	This approximation cannot be used in evaluating
510: \eqsref{iso-max} near $u=0$ because there is cancellation to leading
511: order in all three terms.  In that case, the Taylor expansion,
512: 	$$\eqalign{\erf(u)-u\erf'(u)&={2\over\sqrt\pi}
513: \sum_{k=0}^\infty {(-1)^k\over k!}{2\over 2k+3}u^{2k+3}
514: \cr&={2\over\sqrt\pi}\biggl({2\over 3}u^3 - {2\over 5}u^5+{1\over 7}u^7-{1\over27}
515: u^9+\ldots\biggr),\cr}$$
516: 	may be used.  Alternatively, we can compute \eqsref{iso-max} by
517: numerically evaluating the integrals in \eqsref{iso-gen} with
518: $f_b(v)=f_{bm}(v)$.  This method is then easily extended to include
519: relativistic effects as given in \secref{relativity}.
520: 
521: The ratio $F_{cv}^{a/b}/D_{cvv}^{a/b}$ from \eqsref{iso-max} [and also
522: (\ref{iso-high})] is $-m_av/T_b$ so that the effect of collisions with
523: species $b$ is to make species $a$ approach a Maxwellian with
524: temperature $T_b$.
525: 
526: \subsection{Linearized collision operator}
527: 
528: In many applications in plasma physics (including those involving rf
529: waves) collisions dominate the thermal particles.  Therefore, the
530: distribution function may be expanded about a Maxwellian
531: 	$$f_a({\bf v})=f_{am}(v)+f_{a1}({\bf v}).$$
532: The self-collision operator $C(f_a,f_a)$ may be approximated by the
533: linearized operator
534: 	$$
535: C_{\rm lin}^{a/a}\bigl(f_a({\bf v})\bigr)=
536: 	C\bigl(f_a({\bf v}),f_{am}(v)\bigr)+
537: C\bigl(f_{am}(v),f_a({\bf v})\bigr),
538: \Eqlab{linear}$$
539: 	where we have made use of the fact that $C(f_{am},f_{am})=0$,
540: and we have ignored terms of order $f_{a1}^2$.  We can compute
541: $C\bigl(f_a({\bf v}),f_{am}(v)\bigr)$ using \eqsref{isotropic} and
542: (\ref{iso-max}).  To compute $C\bigl(f_{am}(v),f_a({\bf v})\bigr)$, we
543: express $f_a({\bf v})$ as a sum of Legendre harmonics,
544: \eqref{legend-comp}, to give
545: 	$$C\bigl(f_{am}(v),f_a({\bf v})\bigr)=\sum_{l=0}^\infty
546: C\bigl(f_{am}(v),f_a^{(l)}({\bf v})P_l(\cos\theta)\bigr).
547: \Eqlab{linear-expand}$$
548: 	The zeroth term in the sum can be computed using
549: \eqsref{isotropic} and (\ref{iso-gen}) giving
550: 	$$\eqalignno{{C\bigl(f_{am}(v),f_a^{(0)}(v)\bigr)\over f_{am}(v)}
551: ={4\pi\Gamma^{a/a}\over n_a}\biggl[
552: f_a^{(0)}(v)
553: &{}+\int_0^v {v'{}^2\over v_{ta}^2}f_a^{(0)}(v')
554: \biggl({v'{}^2\over 3v_{ta}^2v}-{1\over v}\biggr)\,dv'\cr
555: &{}+\int_v^\infty {v'{}^2\over v_{ta}^2}f_a^{(0)}(v')
556: \biggl({v^2\over 3v_{ta}^2v'}-{1\over v'}\biggr)\,dv'\biggr].
557: &(\eqlab{iso-lin})\cr}$$
558: 
559: The next term in the sum in \eqref{linear-expand} is
560: given by \eqsref{coll-ros}, (\ref{pot1}), and (\ref{spherical-comp})
561: 	$$\eqalignno{
562: {C\bigl(f_{am}(v), f_a^{(1)}(v)\cos\theta\bigr)\over f_{am}(v)\cos\theta}=
563: {4\pi\Gamma^{a/a}\over n_a}\biggl[f_a^{(1)}(v)
564: &{}+\int_0^v {v'{}^2\over v_{ta}^2}f_a^{(1)}(v')
565: \biggl({v'{}^3\over 5v_{ta}^2v^2}-{v'\over 3v^2}\biggr)\,dv'\cr
566: &{}+\int_v^\infty {v'{}^2\over v_{ta}^2}f_a^{(1)}(v')
567: \biggl({v^3\over 5v_{ta}^2v'{}^2}-{v\over 3v'{}^2}\biggr)\,dv'\biggr].
568: &(\eqlab{first-lin})\cr
569: }$$
570: The corresponding fluxes for this term are
571: $$\eqalignno{{S_v^{a/a}\over f_{am}(v)\cos\theta}=
572: {4\pi\Gamma^{a/a}\over n_a}\biggl[
573: &\int_0^v f_a^{(1)}(v')
574: \biggl({v'{}^5\over 5v_{ta}^2v^3}-{2v'{}^3\over 3v^3}\biggr)\,dv'\cr
575: &+\int_v^\infty f_a^{(1)}(v')
576: \biggl({v^2\over 5v_{ta}^2}+{1\over 3}\biggr)\,dv'\biggr],
577: &(\eqlab{first-flux}a)\cr
578: {S_\theta^{a/a}\over f_{am}(v)\sin\theta}=
579: {4\pi\Gamma^{a/a}\over n_a}\biggl[
580: &\int_0^v f_a^{(1)}(v')
581: \biggl({v'{}^5\over 10v_{ta}^2v^3}-{v'{}^3\over 6v_{ta}^2v}
582: -{v'{}^3\over 3v^3}\biggr)\,dv'\cr
583: &-\int_v^\infty f_a^{(1)}(v')
584: \biggl({v^2\over 15v_{ta}^2}+{1\over
585: 3}\biggr)\,dv'\biggr].&(\ref{first-flux}b)\cr
586: }$$
587: 
588: Because of conservation of number, momentum, and energy, the solutions
589: to the homogeneous equation
590: 	$$
591: C_{\rm lin}^{a/a}\bigl(f_a({\bf v})\bigr)=0$$
592: are
593: 	$$f_a({\bf v})=(a+{\bf b}\cdot m_a{\bf v}+d\half m_av^2)
594: f_{am}(v).$$
595: 	If we substitute $a={\bf b}=0$, we obtain a check on
596: \eqref{iso-lin}. Similarly, $a=d=0$ and ${\bf b}={\bf \hat v}_\parallel$
597: gives a check on \eqref{first-lin}.  Such checks are useful when
598: incorporating the linearized collision operator into a numerical code.
599: 
600: \subsection{Electron-ion collision operator}
601: 
602: We now turn to the specific problem of current drive by lower hybrid
603: waves.  In this problem we wish to solve the Fokker--Planck equation for
604: the electrons including the effects of electron-ion and
605: electron-electron collisions.
606: 
607: Because the ions are so massive relative
608: to the electrons, we have $v\gg v_{ti}$ for nearly all the electrons and
609: \eqsref{iso-high} apply.  Indeed, we can make the further approximations
610: $m_i\rightarrow\infty$, $v_{ti}\rightarrow 0$, in which case the
611: collision operator is given by \eqref{isotropic} with
612: 	$$\eqalignno{D_{cvv}^{e/i}&=F_{cv}^{e/i}=0
613: ,&(\eqlab{ions}a)\cr
614: D_{c\theta\theta}^{e/i}&=
615: \Gamma^{e/e}{Z_i\over 2v},&(\ref{ions}b)\cr}$$
616: 	where
617: 	$$Z_i=-{q_i\ln\Lambda^{e/i}\over q_e\ln\Lambda^{e/e}},$$
618: 	and we have assumed neutrality $q_en_e+q_in_i=0$.  The full
619: electron-ion collision term $C(f_e,f_i)$ can be written as
620: 	$$C^{e/i}\bigl(f_e({\bf v})\bigr)=
621: \Gamma^{e/e}{Z_i\over 2v^3}{1\over\sin\theta}{\partial\over\partial\theta}
622: \sin\theta{\partial\over\partial\theta}f_e({\bf v}).\Eqlab{coll-ions}$$
623: For a
624: multispecies plasma $Z_i$ is replaced by $Z_{\rm eff}$ where
625: 	$$Z_{\rm eff}=
626: {\sum_s n_s q_s^2\ln\Lambda^{e/s}\over n_e q_e^2\ln\Lambda^{e/e}},$$
627: and the sum extends over all the ionic species.
628: 
629: With this collision operator the ions are characterized by a single
630: dimensionless parameter $Z_i$ (or $Z_{\rm eff}$).  The collision
631: operator allows momentum to be transferred from the electrons to the
632: ions, but there is no energy exchange.  The non-negative nature of
633: $f_e$ is preserved.
634: 
635: \subsection{Electron-electron collision operator}
636: 
637: There are several choices for the electron-electron collision
638: operator. We will discuss them starting at the most complex.
639: 
640: The {\sl full} electron-electron collision operator is given by
641: \eqsref{coll-ros} and (\ref{pot1}).  This was first used in
642: current-drive studies by Harvey {\sl et al.}\ \cite{Harvey}.  This
643: collision operator conserves both momentum and energy.  The electron
644: distribution $f_e$ remains non-negative.  Because the collision operator
645: conserves energy, there is nowhere for rf energy absorbed by the
646: electrons to go.  The problem arises because we have reduced a problem
647: in configuration and velocity space to one in velocity space alone, so
648: that there is now no spatial diffusion of energy.  In practice, this
649: problem is solved by inserting an energy loss term into the
650: Fokker--Planck \eqref{fp-1}.  Unfortunately, there are several
651: different models for this loss term and so this procedure is somewhat
652: {\sl ad hoc}.
653: 
654: The {\sl linearized} electron-electron collision operator is
655: given by \eqref{linear}.  This too conserves momentum and energy.
656: The non-negative nature of $f_e$ is no longer preserved.  
657: When the perturbation to $f_{em}$ is small, $f_e$ usually only
658: becomes negative far out on the tail.  The energy conservation of the
659: collision operator again necessitates the introduction of an energy
660: loss term.  Fortunately, there is a systematic way to do this within
661: the context of a Chapman--Enskog--Braginskii expansion
662: \cite{Chapman,Braginskii}.  The energy loss term has the form
663: 	$$-\biggl({m_ev^2\over 2T_e}-{3\over 2}\biggr)f_{em}(v)
664: {\partial\ln T_e\over \partial t},$$
665: 	which appears on the right-hand side of \eqref{fp-1}.
666: Operationally, ${\partial\ln T_e/ \partial t}$ would be adjusted to
667: ensure that the energy of the electron distribution $f_{e}({\bf v})$
668: remained constant.  (In fact, one of the results of the expansion
669: procedure is an equation for the evolution of $T_e$ including the
670: effects of rf and ohmic heating and of energy transport.)
671: 
672: A useful modification of this collision operator is the {\sl truncated}
673: collision operator
674: 	$$C_{\rm trunc}^{e/e}\bigl(f_e({\bf v})\bigr)=
675: C\bigl(f_e({\bf v}),f_{em}(v)\bigr)
676: +C\bigl(f_{em}(v), f_e^{(1)}(v)\cos\theta\bigr),
677: \Eqlab{truncated}$$
678: 	where the first term is given by \eqsref{isotropic} and
679: (\ref{iso-max}) and the second term by \eqref{first-lin}.  This differs
680: from the linearized operator in that we only retain the $l=0$ and $l=1$
681: terms in the sum in \eqref{linear-expand} and we further approximate
682: $f_e^{(0)}(v)$ by $f_{em}(v)$.  As a consequence, this operator
683: conserves momentum but not energy; so there is no need to introduce an
684: energy loss term.  Again, the electron distribution function may become
685: negative.  This collision operator is useful in the study of current
686: drive by low-phase-velocity waves and in the treatment of problems with
687: an electric field.  In both of these examples, a momentum-conserving
688: electron-electron collision operator is required.  This operator was
689: used (in a relativistic form) in the study of current drive by fast
690: waves \cite{relpap}.
691: 
692: A slightly different technique for ensuring momentum conservation  was
693: used in our study of current drive by low-phase-velocity waves
694: \cite{Fisch-Karney}.  There we approximated the electron-electron
695: collision operator by
696: $$C_{\rm drift}^{e/e}\bigl(f_e({\bf v})\bigr)=
697: C\bigl(f_e({\bf v}),f_{em}(\abs{{\bf v}-v_d{\bf \hat v}_\parallel})\bigr),$$
698: 	where the background is a {\sl drifting} Maxwellian with a drift
699: speed $v_d$ adjusted so that the parallel force between $f_e({\bf v})$
700: and the drifting Maxwellian,
701: 	$$P_\parallel^{e/e}=\int m_e S_\parallel^{e/e} \,\dv,
702: \Eqlab{force-def}$$
703: 	vanishes.  This collision operator conserves momentum (by
704: construction) and preserves the non-negative nature of $f_e$.  Energy is
705: not conserved.  It is, however, slightly less accurate than the truncated
706: operator.  In particular, while the truncated operator gives the correct
707: value for the electrical conductivity \cite{Spitzer}, this operator
708: gives an answer which is in error by about $15\%$.  The computation of
709: this collision operator involves computing \eqsref{iso-max} in the
710: drifting frame, converting to cylindrical coordinates using
711: \eqsref{transform}, transforming to the rest frame (which is easy in
712: cylindrical coordinates), and finally converting back to spherical
713: coordinates using \eqsref{transform}.  In order to
714: determine the drift speed, we use the analytical formula for the force
715: on an electron Maxwellian drifting with speed $v_d$ due to a stationary
716: ion background, i.e.,
717: 	$$P_\parallel^{e/i}=-n_e m_e \nu_{te} v_d Z_i
718: {1\over 3}\sqrt{2\over \pi}\sqrt{m_i\over m_i+m_e},$$
719: 	which is valid for $\abs{v_d}\ll v_{te}$.  Here we have taken
720: the mass ratio $m_e/m_i$ to be finite and have assumed that $T_i=T_e$.
721: The force between two electron Maxwellians with a relative drift of
722: $v_d$ is found by taking $Z_i=1$ and $m_i=m_e$ which gives
723: 	$$P_\parallel^{e/e}=-{n_e m_e \nu_{te} v_d
724: \over 3\sqrt\pi}.\Eqlab{force-max}$$
725: 	In the numerical code $P_\parallel^{e/e}$ is computed using
726: \eqref{force-def}.  \Eqref{force-max} is used to estimate the {\sl change} in
727: $v_d$ required to give $P_\parallel^{e/e}=0$.
728: 
729: The situation may be further simplified by assuming that the
730: background electrons are {\sl Maxwellian}, so that the collision
731: operator is given by
732: 	$$C_{\rm Max}^{e/e}\bigl(f_e({\bf v})\bigr)=
733: C\bigl(f_e({\bf v}),f_{em}(v)\bigr),
734: \Eqlab{coll-max}$$
735: 	which may be evaluated using \eqsref{isotropic} and
736: (\ref{iso-max}).  This operator conserves neither energy nor momentum.
737: It does preserve the non-negative nature of $f_e$.  It was used by
738: Kulsrud {\sl et al.}\ \cite{Kulsrud} in the study of runaways, and in
739: studies of lower hybrid current drive \cite{Karney-lh}.  The Maxwellian
740: background serves as a heat bath, so no energy loss terms are required.
741: The absence of momentum conservation introduces approximately a
742: factor-of-two error in the electrical conductivity \cite{Kulsrud} and
743: in the efficiency of current drive by slow waves \cite{Fisch-Karney}.
744: There is a relative error of order $(v_{te}/v)^3$ in the determination
745: of the current-drive efficiency for fast waves \cite{relpap}.
746: 
747: Lastly, $C_{\rm Max}^{e/e}$ may be approximated by using the {\sl high}
748: velocity limit, i.e., by using \eqsref{iso-high} instead of
749: \eqsref{iso-max}.  In fact, because eq.~(\ref{iso-high}b) gives negative
750: diffusion for small $v$, it is usually replaced by
751: 	$$D_{c\theta\theta}^{a/b}=\Gamma^{a/b} {1\over 2v}.$$
752: 	We define the resulting electron-electron collision operator as
753: $C_{\rm high}^{e/e}$.  It has much the same properties as $C_{\rm
754: Max}^{e/e}$.  In particular, it yields a Maxwellian (with
755: temperature $T_e$) as the steady-state solution.  Because of the greater
756: error in the collision term for thermal particles the electrical
757: conductivity is even lower than for $C_{\rm Max}^{e/e}$.  The evaluation
758: of \eqsref{iso-high} is, of course, a little easier to program than that
759: of \eqsref{iso-max}.  However, because the results of evaluating
760: \eqsref{iso-max} can be stored in a table, the extra computational cost of
761: working with $C_{\rm Max}^{e/e}$ is insignificant compared to the
762: solution of the Fokker--Planck equation.  Since $C_{\rm high}^{e/e}$ is
763: less accurate, its use is not recommended for numerical
764: work.  It is, however, useful in analytical work.
765: 
766: When working with these electron-electron collision operators, it is
767: useful to have some benchmark against which to check their numerical
768: realization.  A useful benchmark is provided by the electrical
769: conductivity, which is the ratio of electrical current to electric field
770: in the limit $E\rightarrow 0$.  This is tabulated in \tabref{conduct}
771: for various values of $Z_i$ and for all the electron-electron collision
772: operators discussed here.  These values were obtained by solving the
773: corresponding one-dimensional equation by the method outlined in
774: \secref{adjoint}.  In all cases, the electron-ion collision operator is
775: given by \eqref{coll-ions}.  The conductivity using the full and the truncated
776: electron-electron collision operators is the same as for the
777: linearized operator.
778: In the limit $Z_i\rightarrow\infty$, the conductivity is independent of
779: the electron-electron collision model
780: 	$${J\over E}=16\sqrt{2\over\pi}{1\over Z_i}
781: {n_eq_e^2\over m_e\nu_{te}}.$$
782: 	For the high-velocity approximation to the collision operator
783: the conductivity can be expressed analytically as
784: 	$${J\over E}=
785: {16\over3}\sqrt{2\over\pi}{3Z_i+13\over(Z_i+3)(Z_i+5)}
786: {n_eq_e^2\over m_e\nu_{te}}.$$
787: 
788: \section{Quasilinear Operator}\label{ql}
789: 
790: \subsection{Single wave}
791: 
792: The interaction of electrons (or other species) with a wave is
793: conveniently described in terms of the quasilinear theory
794: \cite{Kennel}.  In this theory the flux of electrons in velocity space is
795: given by
796: 	$${\bf S}_w=-\mat D_w\cdot\nabla f_e,\Eqlab{wave-flux}$$
797: 	where $\mat D_w$ is the quasilinear diffusion tensor which
798: depends on the waves present in the plasma.  Although quasilinear theory
799: is not strictly applicable to a single wave, we will start with this
800: case because it is the simplest.  Suppose there is a uniform wave
801: present in the plasma, i.e., \def\re{\mathop{\rm Re}\nolimits}%
802: 	$${\bf E}({\bf r},t)=\re[{\bf E}_w\exp(i{\bf k\cdot r}-i\omega t)].
803: \Eqlab{single-wave}$$
804: 	The quasilinear diffusion coefficient is given by \cite{Kennel}
805: 	$$\mat D_w=\sum_n{\pi\over 2}{q_e^2\over m_e^2}
806: \delta(\omega-k_\parallel v_\parallel-n\Omega_e){\bf a}_n^\ast{\bf a}_n
807: \Eqlab{ql-single}$$
808: and
809: $$\eqalignno{
810: {\bf a}_n&=\Theta_n{k_\parallel\over \omega}\biggl[\biggl(
811: {\omega\over k_\parallel}-v_\parallel\biggr){\bf \hat v}_\perp+
812: v_\perp{\bf \hat v}_\parallel\biggl]\cr
813: \Theta_n&={E_{w+} J_{n-1} + E_{w-} J_{n+1}\over \sqrt 2}+
814: {v_\parallel\over v_\perp}J_nE_{w\parallel},\cr}$$
815: 	where $\Omega_e=q_eB/m_e$ is the electron cyclotron frequency,
816: $B$ is the magnetic field, $\ast$ indicates complex conjugation,
817:  $J_n$ is the $n$th order Bessel function,
818: and the argument of the Bessel functions is $k_\perp v_\perp/\Omega_e$.
819: $E_{w+}$ and $E_{w-}$ are the left- and right-handed components of
820: ${\bf E}_w$; in a right-handed
821: cartesian coordinate system with $\bf \hat z$ parallel to $\bf B$ and
822: $\bf k$ lying in the $(x,z)$ plane, we
823: have
824: 	$$\eqalign{E_{w+}&={E_{wx}+iE_{wy}\over\sqrt2},\cr
825: E_{w-}&={E_{wx}-iE_{wy}\over\sqrt2},\cr
826: E_{w\parallel}&=E_{wz}.\cr}$$
827: 
828: It is instructive to consider the properties of \eqref{ql-single}.  The
829: delta function specifies the resonance condition.  Only particles for
830: which the Doppler-shifted wave frequency $\omega-k_\parallel
831: v_\parallel$ is zero ($n=0$---the Landau resonance) or a multiple of
832: the cyclotron frequency ($n\ne 0$---a cyclotron harmonic resonance)
833: interact with the wave.  The vector ${\bf a}_n$ is perpendicular to the
834: velocity of the electron in the wave frame ${\bf
835: v}-(\omega/k_\parallel){\bf \hat v}_\parallel$.  This means that the
836: wave-induced flux is along diffusion paths which lie in constant-energy
837: surfaces in the wave frame; see \figref{diff-fig}.  Similarly, the flux
838: is proportional to the gradient in $f_e$ in this direction.  As a
839: consequence, when an electron interacts with a particle via the Landau
840: resonance, the diffusion tensor consists of only a single component
841: 	$$\mat D_w=D_{\parallel\parallel}
842: {\bf \hat v}_\parallel{\bf \hat v}_\parallel.$$
843: Likewise, for a cyclotron harmonic resonance, we have
844: $$\mat D_w=D_{\perp\perp}
845: {\bf \hat v}_\perp{\bf \hat v}_\perp,$$
846: provided that $v_\perp$ is small compared with $n\Omega_e/k_\parallel$.
847: 
848: The appearance of the delta function in \eqref{ql-single} is a
849: consequence of the assumed uniformity of the magnetic field.  In this
850: case, $v_\parallel$ is a constant of the unperturbed motion and so a
851: particle remains in resonance for a long time.  In situations described
852: by bounce-averaged codes, the magnetic field and $v_\parallel$ vary
853: along a particle orbit so that the particle does not remain in
854: resonance.  This effect can be taken into account by averaging
855: \eqref{ql-single} along a particle trajectory \cite{Bernstein}.  This
856: removes the delta function, although there are still singularities in
857: the resulting expression arising from those particles which turn in the
858: resonance \cite{Kerbel}.
859: 
860: \subsection{Many waves}
861: 
862: \Eqref{ql-single} is easily generalized to include a more realistic
863: representation of the wave fields.  An important application is to the
864: incorporation of quasilinear effects into a ray-tracing code.  Here the
865: externally injected rf power is represented by several rays.  Let us
866: consider the interaction of these waves with the electrons on a
867: particular flux surface.  At the point where a given ray intersects the
868: flux surface it is characterized by its position $\bf r$, wave number
869: $\bf k$, and power $W$ (usually the frequency $\omega$ is fixed by the
870: rf source).  $W$ measures the number of watts carried by the ray.  In
871: order to apply \eqref{ql-single}, we must determine the amplitude of
872: the corresponding single wave ${\bf E}_w$ which has the same
873: polarization and same rms field amplitude as the ray (with the rms
874: averaging performed over time and over the flux surface).
875: 
876: The ray contributes
877: 	$$U={W\over \abs{{\bf v}\!_g\cdot {\bf\hat n}}A_f}$$
878: 	to the wave energy density (in $\rm J/m^3$) averaged over the
879: flux surface, where ${\bf v}\!_g$ is the group velocity of the ray, $\bf
880: \hat n$ is the unit vector normal to surface at the point of
881: intersection and $A_f$ is the area of the flux surface.  The polarization of
882: the electric field is given by
883: 	$$\mat K\cdot {\bf E}_w=0,$$
884: 	where
885: 	$$\mat K={c^2\over \omega^2}({\bf kk}-k^2\mat I)
886: +\mat I+i{\bsigma({\bf k},\omega)\over \omega\epsilon_0}$$
887: 	is the dispersion tensor, $c$ is the velocity of light, and
888: $\bsigma({\bf k},\omega)$ is the conductivity tensor.  The energy
889: density is related to ${\bf E}_w$ by \cite{Bers}
890: 	$$U={1\over 4}\epsilon_0\omega {\bf E}_w^*\cdot
891: {\partial \mat K\over \partial \omega}\cdot {\bf E}_w.$$
892: 	Given $W$, we can therefore determine ${\bf E}_w$ (to within an
893: ignorable phase factor) appropriately averaged over the flux surface.
894: 
895: This is substituted into \eqref{ql-single} and the result summed over
896: all the rays to give the overall quasilinear diffusion tensor.  In
897: practice, the delta-functions appearing in this expression must be
898: replaced by smoothed functions.  This allows the ray-tracing procedure
899: to reflect the true situation in which a continuous spectrum of waves
900: is launched.
901: 
902: We complete the discussion of the ray-tracing by pointing out that the
903: damping of the rays should be calculated self-consistently from $\mat
904: D_w$.  The power that a particular ray loses per unit volume due to
905: absorption by the electrons is given by \eqref{pd-def}, where instead
906: of the total ${\bf S}_w$ we use the contribution the ray in question
907: makes to ${\bf S}_w$.  To this should be added the power absorbed by the
908: other species if applicable.  Then the ray power $W$ satisfies the
909: equation
910: 	$${dW\over dt}=-P\abs{{\bf v}\!_g\cdot {\bf\hat n}}A_f,$$
911: where the time derivative is the derivative taken along the ray.
912: 
913: If instead of a discrete set of waves, the wave fields are given by a
914: spectrum
915: 	$${\bf E}({\bf r},t)=
916: \int {\bf E}_w({\bf k})\exp[i{\bf k\cdot r}-i\omega({\bf k}) t]\,
917: {d^3{\bf k}\over (2\pi)^3},$$
918: then \eqref{ql-single} becomes \cite{Kennel}
919: 	$$\mat D_w=\sum_n{q_e^2\over m_e^2}
920: \int\!{d^3{\bf k}\over (2\pi)^3}{1\over V_p}
921: \pi\delta[\omega({\bf k})-k_\parallel v_\parallel-n\Omega_e]
922: {\bf a}_n^\ast{\bf a}_n,
923: \Eqlab{ql-spectrum}$$
924: 	where $V_p$ is the configuration space volume of the plasma and
925: the definition of ${\bf a}_n$ is generalized in the obvious way.
926: 
927: \subsection{Model forms}
928: 
929: The results given above allow a ray-tracing code to be coupled to the
930: solution of the Fokker--Planck equation.  This is an extremely
931: complicated system, and much work has been carried out using assumed
932: forms for the quasilinear diffusion coefficient.  This allows us to
933: study the physics of the interaction of the electrons and the waves
934: without having to worry about the additional physics of the wave
935: propagation.  The most widely used model form for lower hybrid waves was
936: introduced by Fisch \cite{Fisch1} and is given by
937: 	$$\mat D_w=D_w(v_\parallel) {\bf \hat v}_\parallel{\bf \hat v}_\parallel,
938: \eqn(\eqlab{d-ql}a)$$
939: where
940: $$D_w(v_\parallel)=\cases{D_0,&for $v_1<v_\parallel<v_2$,\cr
941: 0,&otherwise.\cr}\eqn(\ref{d-ql}b)$$
942: 	This form of $\mat D_w$ is justified as follows.  Because
943: lower hybrid waves interact only via the Landau resonance, only the
944: ${\bf \hat v}_\parallel{\bf \hat v}_\parallel$ component is present.  If
945: $k_\perp v_{te}/\Omega_e\ll 1$, the dependence on perpendicular
946: velocity may be ignored ($J_0\approx 1$).  Finally, in many cases, the
947: magnitude of the quasilinear diffusion greatly dominates over the
948: collisions; thus the quasilinear diffusion coefficient tends to make an
949: abrupt transition (in velocity space) from being negligible to being
950: large; if $D_0$ is sufficiently large (i.e., large enough to form a
951: quasilinear plateau), this situation is accurately modeled by
952: eq.~(\ref{d-ql}b).
953: 
954: This particular form for $\mat D_w$ is useful because much theoretical
955: work has been carried out using it \cite{Fisch1}.  Numerical solutions
956: to the Fokker--Planck equation provide the best test of these
957: theories. It is therefore important that any numerical code be able to
958: handle the discontinuities in $\mat D_w$.  (Note, however, that both
959: $f_e$ and $\bf S$ are continuous even if $\mat D_w$ is not.)
960: 
961: This model is readily generalized, for example, by allowing
962: $D_0(v_\parallel)$ to be an arbitrary function.  Thus the effect of a
963: backward component to the lower hybrid spectrum can be studied by
964: including another boxlike component to $D_0$.  Similar models have
965: been used to study low-phase-velocity current drive \cite{Fisch-Karney}
966: and electron-cyclotron current drive \cite{Karney-ec}.
967: 
968: \subsection{Direct specification of the quasilinear flux}
969: 
970: Both analytical and numerical studies show that the current drive
971: efficiency is primarily determined by the location at which electrons
972: interact with the waves and the direction in which the waves push the
973: electrons.  It is sometimes useful to specify the rf-induced flux
974: directly as some arbitrary vector field ${\bf S}_w({\bf v})$.  Indeed,
975: in some cases we may know ${\bf S}_w$ more accurately than we know
976: $\mat D_w$.  In a ray-tracing calculation, $\mat D_w$ may be calculated
977: self-consistently in terms of the power flows in the various rays.
978: However, in cruder zero-dimensional calculations, we may wish to assert
979: merely that so much rf power is absorbed by the electrons.  Then ${\bf
980: S}_w$ may be estimated from \eqref{pd-def} using an {\sl a priori}
981: knowledge of which electrons interact with the waves.  Alternatively,
982: ${\bf S}_w$ may be estimated from either an approximate analytic
983: solution of the Fokker--Planck equation \cite{asymp} or from a solution
984: of the one-dimensional Fokker-Planck equation \cite{Fisch1}.
985: 
986: If ${\bf S}_w$ is given, then the Fokker--Planck equation (\ref{fp-1}) is
987: an inhomogeneous (instead of homogeneous) equation.  However, assuming
988: that one of the linear electron-electron collision operators is being
989: used, the linear operator acting on $f_e$ in \eqref{fp-1} is now
990: independent of the wave drive.  This property is used in the adjoint
991: methods to provide a very efficient method of solving for moments of
992: $f_e$ (see \secref{adjoint}).
993: 
994: \section{Boundary Conditions}\label{bcs}
995: 
996: \subsection{Computational domain}
997: 
998: We shall take the computational domain $V$ for the Fokker--Planck equation
999: to be
1000: $$0<v_\perp<v_{\perp\rm max},\qquad
1001: v_{\parallel\rm min}<v_\parallel<v_{\parallel\rm max},
1002: \Eqlab{cyl-domain}$$
1003: 	for problems solved in a cylindrical coordinate system and
1004: $$0<v<v_{\rm max},\qquad 0<\theta<\pi,\Eqlab{spher-domain}$$
1005: 	for problems solved in a spherical coordinate system.  The
1006: boundary of $V$ is defined to be $A$.  (For example, in spherical
1007: coordinates, $A$ is the spherical surface $v=v_{\rm max}$.)
1008: 
1009: \subsection{Internal boundaries}
1010: 
1011: We distinguish two types of boundary:  internal and external boundaries.
1012: The internal boundaries are the simplest.  In a cylindrical
1013: coordinate system we have an internal boundary at $v_\perp=0$.  Values
1014: of $f$ beyond this boundary are determined by symmetry
1015: 	$$f_e(-v_\perp,v_\parallel)=f_e(v_\perp,v_\parallel).\Eqlab{cyl-bc}$$
1016: Similarly, in spherical coordinates we have internal boundaries at $v=0$
1017: and at $\theta=0$ and $\theta=\pi$.  These boundaries are treated with
1018: the boundary conditions
1019: 	$$\eqalignno{
1020: f_e(-v,\theta)&=f_e(v,\pi-\theta),&(\eqlab{spher-bc}a)\cr
1021: f_e(v,-\theta)&=f_e(v,\theta),&(\ref{spher-bc}b)\cr
1022: f_e(v,\pi+\theta)&=f_e(v,\pi-\theta).&(\ref{spher-bc}c)\cr}$$
1023: 
1024: \subsection{External boundaries}
1025: 
1026: The other boundaries are inserted into the problem in violation of the
1027: true physical picture.  In reality the velocity domain extends off to
1028: infinity; on the computer, however, we normally study only a subspace.
1029: We have to choose the subspace to include all the interesting physics:
1030: for studies of electron distribution in a spherical coordinate system,
1031: we require $v_{\rm max}\gg v_{te}$; if the electrons are driven by
1032: lower hybrid waves, then we further require $v_{\rm max}>
1033: (\omega/k_\parallel)_{\rm max}$, the maximum wave phase velocity; if we
1034: wish to study runaways, then $v_{\rm max}$ must exceed the runaway
1035: velocity; and so on.  We next have to choose boundary conditions which
1036: are as ``innocuous'' as possible; i.e., which perturb the solution in
1037: the domain of integration as little as possible compared to the
1038: solution in the full domain.
1039: 
1040: For electron current-drive problems we choose the condition
1041: 	$${\bf S\cdot \hat n}=0,\Eqlab{bc-a}$$
1042: 	on the external boundary $A$, where $\bf\hat n$ is the normal
1043: to $A$.  This means that plasma cannot enter or leave the domain of
1044: integration.  Thus the number of electrons is conserved with this
1045: boundary condition.  This boundary condition gives a Maxwellian
1046: steady state in the absence of the rf, and allows a steady-state
1047: solution to be reached in the presence of rf.
1048: 
1049: If an electric field is present, then in the real problem some
1050: electrons will run away.  Now we wish to impose boundary conditions
1051: which ``allow'' this to happen.  At the boundary we have $v\gg v_{te}$
1052: so that collisions are weak, and the dominant process is the
1053: acceleration by the electric field (we assume that the boundary is
1054: removed from the region where the rf diffusion takes place).  
1055: The Fokker--Planck equation then reduces to a hyperbolic equation.  The
1056: tactic is to apply the same boundary condition as before, namely
1057: \eqref{bc-a}, where the characteristics of the hyperbolic system enter
1058: the domain of integration.  Where the characteristics leave, we set
1059: those diffusion terms which lead to a flux across the boundary to zero.
1060: This makes the equation purely hyperbolic in the direction normal to
1061: the boundary and so {\sl no} boundary condition is required.  (We shall
1062: see in \secref{spatial} how this comes about in the numerical scheme.)
1063: 
1064: If we assume that $q_eE>0$ so that electrons run away in the positive
1065: direction, then in cylindrical coordinates we would impose
1066: 	$$\eqalign{S_\parallel=0,\qquad
1067: &\hbox{for }v_\parallel=v_{\parallel\rm min},\cr
1068: S_\perp=0,\qquad
1069: &\hbox{for }v_\perp=v_{\perp\rm max},\cr
1070: D_{\parallel\perp}=D_{\parallel\parallel}=0,\qquad
1071: &\hbox{for }v_\parallel=v_{\parallel\rm max}.\cr}\Eqlab{bc-e-cyl}$$
1072: (The boundary at $v_\perp=v_{\perp\rm max}$ is taken to be an incoming
1073: boundary because the small collisional friction makes the
1074: characteristics enter along this boundary.)
1075: 
1076: A slightly more accurate treatment is possible in spherical
1077: coordinates.  If we compare the various collision terms in the
1078: high-velocity limit \eqsref{iso-high}, we find
1079: $F_{cv}\sim D_{c\theta\theta}/v\sim 1/v^2$ and
1080: $D_{cvv}/v\sim v_{te}^2/v^4$.  Thus we can ignore the energy diffusion
1081: term $D_{cvv}$ compared with the other collisional terms.  The
1082: pitch-angle scattering term $D_{c\theta\theta}$ requires no special
1083: handling because it causes diffusion parallel to the boundary.  The
1084: equation is, therefore, hyperbolic in the direction perpendicular to the
1085: boundary with a characteristic acceleration
1086: given by $F_v=F_{cv}+(q_e E/m_e)\cos\theta$.  The boundary conditions
1087: on $v=v_{\rm max}$ then become
1088: 	$$\eqalign{S_v=0,\qquad
1089: &\hbox{for }F_{v}<0,\cr
1090: D_{vv}=D_{v\theta}=0,\qquad
1091: &\hbox{for }F_{v}>0.\cr}\Eqlab{bc-e-spher}$$
1092: 	For $v_{\rm max}\gg v_{te}$, $F_{cv}$ is accurately
1093: approximated by eq.~(\ref{iso-high}c) (with $a=b=e$).  Thus, if $\abs
1094: E<m_e\Gamma^{e/e}/\abs{q_e}v_{\rm max}^2$, this boundary condition
1095: reduces to \eqref{bc-a}, allowing problems involving both an electric
1096: field and rf diffusion to be handled in a unified way.  In this small
1097: electric field limit, $S_v=0$ is zero everywhere on the boundary and
1098: the numerical runaway rate vanishes.  This is a close approximation to
1099: the true situation in which the runaway rate is exponentially
1100: small---on the order of $\exp(-v_{\rm max}^2/v_{te}^2)$.
1101: 
1102: \subsection{Treatment of runaways}
1103: 
1104: With a finite boundary, we can determine the runaway rate accurately
1105: (provided $v_{\rm max}$ is sufficiently large).  However, the behavior
1106: of the runaways beyond the boundary is not followed.  One could, of
1107: course, just choose a very large boundary; but this is wasteful of
1108: computer resources and really just postpones the time at which the
1109: problem is encountered.  It is, therefore, preferable to treat the
1110: runaways as a separate species.  Assuming that the runaways are affected
1111: only by the electric field, the density and current moments of the
1112: runaway population form a closed set of equations.  We define
1113: 	$$\eqalign{n_r&=\int_{\overline V} f_e\,\dv,\cr
1114: J_r&=\int_{\overline V} q_ev_\parallel f_e\,\dv,\cr}$$
1115: 	where $\overline V$ is the complement of $V$, i.e., the region
1116: $v>v_{\rm max}$ in spherical coordinates.  Applying eqs.~(\ref{cons}a)
1117: and (\ref{cons}b) to $\overline V$ we find
1118: 	$$\eqalignno{
1119: {\partial n_r\over \partial t}&=
1120: \int_A{\bf S}\cdot\da,\cr
1121: {\partial J_r\over \partial t}&=
1122: {q_e^2 E\over m_e} n_r+\int_A q_ev_\parallel {\bf S}\cdot\da.\cr}$$
1123: 	Thus if we wish to determine the total current as a function of
1124: time, we need only supplement the Fokker--Planck equation by two
1125: ordinary differential equations and then sum the nonrunaway and
1126: runaway contributions to the current.
1127: 
1128: \section{Spatial Differencing}\label{spatial}
1129: \subsection{Choice of coordinate system}
1130: 
1131: We have discussed both the cylindrical and the spherical coordinate
1132: systems.  Which one should be used in a given application?  The numerical
1133: scheme that is described here works best if the diffusion tensor is
1134: nearly diagonal.  Then the mixed derivative terms in \eqsref{cylindrical}
1135: or (\ref{spherical}) are small.  (It is these terms which tend to make
1136: the numerical scheme unstable.)  Now the collision operator is
1137: approximately diagonal in spherical coordinates while the quasilinear
1138: term is nearly diagonal in cylindrical coordinates.  Thus the choice of
1139: coordinate system to some extent depends on the relative strength of
1140: these two terms.  Cylindrical coordinates were used in the study of current
1141: drive by low-phase-velocity waves \cite{Fisch-Karney} because the edges
1142: of the resonant region line up with coordinate lines allowing the
1143: scaling with phase velocity to be measured more accurately.  On the
1144: whole, however, the spherical system is to be preferred because the electron-ion
1145: collision term \eqref{coll-ions} becomes large near $v=0$ and we wish this
1146: term to be diagonal.  In \refref{Fisch-Karney} much smaller time
1147: steps had to be taken to avoid the problem with the electron-ion term.
1148: The boundary conditions can also be applied more accurately in
1149: spherical coordinates when an electric field is present
1150: [\eqsref{bc-e-spher}].  For this reason, we will focus on the spherical
1151: coordinate system in this section.  Extension to the cylindrical
1152: coordinate system is straightforward.
1153: 
1154: An alternate representation of $f_e$ is as a series of Legendre
1155: harmonics.  This has no particular merit in quasilinear problems
1156: because the sharp gradients in $\mat D_w$, \eqref{d-ql}, cause the
1157: Legendre expansion to be slowly convergent.
1158: 
1159: \subsection{Normalizations}
1160: 
1161: In solving equations of physical significance on the computer, it is
1162: often useful to normalize all the physical quantities.  This allows us
1163: to work with numbers which are closer to unity (and thus avoid
1164: potential problems due to arithmetic overflow or underflow); more
1165: importantly, the number of parameters needed to specify the problem is
1166: often reduced.
1167: 
1168: For the problem of current drive by lower hybrid waves, we solve the
1169: Fokker--Planck equation for the electrons.  We normalize velocities to
1170: $v_{te}$ \eqref{vts-max}, times to $\tau_{te}$ \eqref{nute}, the electron
1171: density to $n_e$, the electron distribution to $n_e/v_{te}^3$, the
1172: quasilinear diffusion coefficient to $v_{te}^2\nu_{te}$, the electric
1173: field to $m_e v_{te}\nu_{te}/q_e$, the current density to
1174: $n_eq_ev_{te}$, power density to $n_em_ev_{te}^2\nu_{te}$, etc.
1175: 
1176: These normalizations coincide with those used by Kulsrud {\sl et al.}\
1177: \cite{Kulsrud}.  However, they differ from those used in some of our
1178: earlier papers, e.g., \refref{Karney-lh}.  (The thermal collision time
1179: differs by a factor of two.)
1180: 
1181: Since we are only dealing with the electron distribution, we will
1182: drop the species label from $f$ and other electron quantities.
1183: Otherwise, we shall use the same notation for normalized and
1184: unnormalized quantities.  For example, the electron Maxwellian
1185: \eqref{max} reads in normalized terms
1186: 	$$f_m(v)={1\over (2\pi)^{3/2}}\exp(-\half v^2).$$
1187: 
1188: The reduction in the number of parameters now becomes apparent.  The
1189: plasma is characterized by a single parameter $Z_i$ and the quasilinear
1190: diffusion coefficient by three parameters $D_0$, $v_1$, and $v_2$.
1191: 
1192: \subsection{The numerical grid}
1193: 
1194: We wish to solve \eqref{fp-1} in the domain $V$ \eqref{spher-domain}.
1195: We do this by converting the differential equation to an algebraic
1196: equation using the finite difference method.  In this method $f$ is
1197: represented by its values on finite set of points and differentials are
1198: represented by differences between neighboring values.
1199: 
1200: First, we establish a numerical grid by
1201: dividing $v$ and $\theta$ into $N$ and $M$ equal pieces, respectively.
1202: Thus we define
1203: 	$$\Delta v=v_{\rm max}/N,\qquad \Delta\theta=\pi/M,\Eqlab{delta-def}$$
1204: 	together with grid  positions
1205: 	$$\eqalignno{v_j&=j\,\Delta v,&(\eqlab{vj}a)\cr
1206: 	\theta_i&=i\,\Delta\theta.&(\ref{vj}b)\cr}$$
1207: 	This grid system defines a system of cells.  The electron
1208: distribution function is represented by its values at the {\sl centers}
1209: of these cells, i.e., by the values
1210: 	$$f_{i+1/2,j+1/2}=f(v_{j+1/2},\theta_{i+1/2}),
1211: 	\quad\hbox{for } 0\le i<M,\quad 0\le j<N,$$
1212: 	with $i$ and $j$ being integers; see \figref{grid-fig}.  The
1213: cell $v_j<v<v_{j+1}$, $\theta_i<\theta<\theta_{i+1}$ ($i$ and $j$
1214: integers) is assigned a volume
1215: 	$$V_{i+1/2,j+1/2}=2\pi\sin\theta_{i+1/2}v_{j+1/2}^2\,
1216: \Delta v\Delta\theta. \Eqlab{vol}$$
1217: We will define numerical volume integration by
1218: \def\numint{\mathop{\rm int}\nolimits}%
1219: \def\flux{\mathop{\rm flux}\nolimits}%
1220: $$\numint{X}=\sum_{i=0}^{M-1}\sum_{j=0}^{N-1}
1221: X_{i+1/2,j+1/2}f_{i+1/2,j+1/2}V_{i+1/2,j+1/2}.\Eqlab{num-int}$$
1222: 	This is the discrete analogue of $\int_V Xf \,\dv$; see
1223: \eqref{int-spher}.  We define the flux of a quantity through
1224: the boundary by
1225: 	$$\flux X=\sum_{i=0}^{M-1}
1226: 2\pi\sin\theta_{i+1/2} v_N^2 X_{i+1/2,N+1/2}
1227: S_{v,i+1/2,N}\,\Delta\theta,\Eqlab{cons-loss}$$
1228: 	which is a discrete analogue of $\int_A X{\bf S}\cdot\da$.  The
1229: number density of electrons becomes
1230: 	$$n=\numint1.\Eqlab{n-grid}$$
1231: 
1232: An alternative approach to finite differences is provided by the
1233: finite-element method where the $f$ is represented by the superposition
1234: of a set of trial functions with finite support.  This approach has been
1235: used in Fokker--Planck codes by workers at Lausanne \cite{Kritz,Succi}.
1236: The finite-element method is also used in some commercial computer
1237: codes for the solution of partial differential equations.  One such code
1238: has been applied to the Fokker--Planck equation by Fuchs {\sl et al.}\
1239: \cite{Fuchs}.  If we identify the weights of the trial functions with
1240: the values of $f$ at the grid positions, we see that the
1241: finite-difference and finite-element methods are quite similar.  In
1242: particular, the goals of the methods are identical: to express
1243: algebraically $\partial f/\partial t$ at a particular location in terms
1244: of $f$ at the same and neighboring locations (usually, the eight
1245: nearest neighbors).  Thus our discussion of the time advancement of the
1246: equation in \secref{time} is independent of the choice of method.
1247: 
1248: \subsection{Divergence of flux}
1249: 
1250: Consider the Fokker--Planck equation in the form \eqref{fp-2}.
1251: This is translated onto our numerical grid in a conservative form as
1252: 	$$\eqalignno{{\partial f_{i+1/2,j+1/2}\over \partial t}=-\biggl(&
1253: {v_{j+1}^2 S_{v,i+1/2,j+1}-v_j^2 S_{v,i+1/2,j}\over v_{j+1/2}^2 \,\Delta v}\cr
1254: &+
1255: {\sin\theta_{i+1} S_{\theta,i+1,j+1/2}-\sin\theta_i
1256: S_{\theta,i,j+1/2}\over v_{j+1/2}\sin\theta_{i+1/2}\,\Delta \theta}\biggr).
1257: &(\eqlab{flux-grid})}$$
1258: 	Notice that the fluxes are required on the edges of the cells
1259: (see \figref{grid-fig}) and that the fluxes on the internal boundaries
1260: do not contribute since they are multiplied by $v_0=0$ or $\sin
1261: \theta_0=\sin\theta_M=0$.  With this method we difference the fluxes
1262: and not the diffusion and friction coefficients.  This lets us treat
1263: problems in which $\mat D_w$ is discontinuous, e.g., as given by
1264: \eqsref{d-ql}.  The scheme in \eqref{flux-grid} is accurate to second
1265: order in $\Delta v$ and $\Delta \theta$.
1266: 
1267: This form of difference equation is called conservative because it
1268: obeys the conservation law
1269: 	$${\partial \numint 1\over \partial t}+\flux 1=0,
1270: \Eqlab{cons-grid}$$
1271: 	where $\numint$ and $\flux$ are defined by \eqsref{num-int} and
1272: (\ref{cons-loss}).  This is a discrete counterpart of eq.~(\ref{cons}a).
1273: If $S_{v,i+1/2,N}=0$ for all $i$, then we have $\flux 1=0$ and particles are
1274: exactly conserved in the numerical scheme (if we ignore round-off
1275: errors).  The discrete form of the parallel component of the
1276: momentum conservation law eq.~(\ref{cons}b) is
1277: 	$$\eqalignno{{\partial \numint (v\cos\theta)\over \partial t}
1278: +\flux (v\cos\theta)
1279: &=\sum_{i=0}^{M-1}\sum_{j=0}^N
1280: 2\pi v_j^2\sin\theta_{i+1/2}\cos\theta_{i+1/2} S_{v,i+1/2,j}
1281: \,\Delta v\Delta\theta\qquad\cr
1282: &\quad{}-\sum_{i=0}^{M}\sum_{j=0}^{N-1}
1283: 2\pi v_{j+1/2}^2\sin^2\!\theta_{i}S_{\theta,i,j+1/2}
1284: \,\Delta v\,2\sin(\half\Delta\theta),&(\eqlab{momcons-grid})\cr}$$
1285: while the energy conservation relation eq.~(\ref{cons}c) becomes
1286: 	$${\partial \numint(\half v^2)\over\partial t}
1287: +\flux(\half v^2)=
1288: \sum_{i=0}^{M-1}\sum_{j=0}^{N}
1289: 2\pi\sin\theta_{i+1/2} v_j^3 S_{v,i+1/2,j} \,\Delta v\Delta\theta.
1290: \Eqlab{encons-grid}$$
1291: 	These relations are useful in that they establish definitions
1292: of various physical quantities that are consistent with the numerical
1293: scheme.  For example, we can interpret the right-hand side of
1294: \eqref{encons-grid} as the total power flowing into the electrons.  This
1295: definition is consistent with the numerical definition of the energy of
1296: the electrons, namely $\numint(\half v^2)$.  Furthermore, we can
1297: determine the power flowing into the electrons from the waves (for
1298: example) by replacing $S_v$ in the right-hand side of this equation by
1299: the flux due to the waves $S_{wv}$ [compare with
1300: \eqref{pd-def}].  In this way, we obtain a complete and accurate power
1301: balance for the electrons.  Similarly, the right-hand side of
1302: \eqref{momcons-grid} gives the definition of the force on the
1303: electrons.  This is used when evaluating $P_\parallel^{e/e}$ in
1304: \eqref{force-def}.
1305: 
1306: [In order to prove \eqsref{momcons-grid} and (\ref{encons-grid}), the
1307: following relation is useful:
1308: 	$$\sum_{i=0}^{M-1}
1309: \half(A_{i+1}+A_i)(B_{i+1}-B_i)
1310: =A_MB_M-A_0B_0-\sum_{i=0}^{M-1}
1311: \half(B_{i+1}+B_i)(A_{i+1}-A_i).$$
1312: This is the rule for  ``summing by parts''---the discrete counterpart
1313: of integration by parts.]
1314: 
1315: The basic difference equation (\ref{flux-grid}) is readily generalized
1316: to nonuniform grids.  However, the derivation of \eqsref{momcons-grid}
1317: and (\ref{encons-grid}) relies on the uniformity of the grid and they
1318: cannot easily be generalized.  Nonuniform spacing is used in
1319: {\sc FPPAC} \cite{McCoy}.
1320: 
1321: \subsection{Stream function}
1322: 
1323: A very useful tool for understanding the Fokker--Planck equation
1324: (\ref{fp-2}) is the flux plot, which shows the vector field ${\bf
1325: S}({\bf v})$.  This is sometimes displayed as a set of arrows, one at
1326: each grid point, which point in the direction of ${\bf S}$ and which
1327: have a length proportional to $S$.  In this problem, $S_v$ and
1328: $S_\theta$ are known at different locations, so that realization of
1329: this prescription would necessitate interpolation.  Furthermore, such a
1330: display is often very misleading because the visual impression is
1331: strongly affected by whether the arrows line up with other grid points
1332: or not---a purely artificial aspect of the problem.
1333: 
1334: The much superior method is possible if we restrict ourselves to the
1335: steady state.  In this case, the vector field ${\bf S}({\bf v})$ is
1336: divergence-free $\nabla\cdot{\bf S}=0$, and so may be expressed as the
1337: curl of a stream function, i.e.,
1338: 	$${\bf S}({\bf v})=\nabla\times{nA({\bf v}){\bf \hat{\bphi}}
1339: \over 2\pi v\sin\theta},$$
1340: 	where $\phi$ is the azimuthal coordinate.  The components of
1341: $\bf S$ are given by
1342: 	$$\eqalignno{
1343: S_v&={n\over 2\pi v^2\sin\theta}{\partial A\over\partial \theta},
1344: &(\eqlab{stream}a)\cr
1345: S_\theta&=-{n\over 2\pi v\sin\theta}{\partial A\over\partial v}.
1346: &(\ref{stream}b)\cr}$$
1347: 	Because ${\bf S}\cdot\nabla A=0$, lines of constant $A$ are
1348: stream lines.  Thus a contour plot of $A({\bf v})$ gives the vector
1349: field of ${\bf S}({\bf v})$.  The stream lines are obviously closed
1350: (indicating that the flow is divergence-free), and the total flux of
1351: electrons between any two contours is equal to the difference in the
1352: values of $nA$ on those two contours.
1353: 
1354: We can compute $A$ on the numerical grid using discrete analogs of
1355: \eqsref{stream}
1356: 	$$\eqalignno{A_{i,j}&={2\pi v_j^2\over n}\sum_{i'=0}^{i-1}
1357: \sin\theta_{i'+1/2}S_{v,i'+1/2,j}\,\Delta\theta,
1358: &(\eqlab{stream-grid}a)\cr
1359: &=-{2\pi\sin\theta_i\over n}\sum_{j'=0}^{j-1}
1360: v_{j'+1/2}S_{\theta,i,j'+1/2}\,\Delta v.
1361: &(\ref{stream-grid}b)\cr}$$
1362: 	If $\partial f_{i+1/2,j+1/2}/\partial t=0$ according to
1363: \eqref{flux-grid}, then these two definitions are consistent.
1364: 
1365: \subsection{Computation of the flux}
1366: 
1367: In order to complete the specification of the difference scheme we must
1368: give formulas for $S_{v,i+1/2,j}$ and $S_{\theta,i,j+1/2}$ in
1369: \eqref{flux-grid}.  These depend on the type of electron-electron
1370: collision operator used.  We start with collisions off a
1371: Maxwellian background $C_{\rm Max}^{e/e}$, \eqref{coll-max}.  This is
1372: the simplest case and yet it exhibits all the difficulties of solving
1373: the Fokker--Planck equation.
1374: 
1375: The collisional flux is given by the sum of the flux contributing to
1376: $C_{\rm Max}^{e/e}$ which is given by \eqsref{isotropic} and
1377: (\ref{iso-max}) and the flux contributing to $C^{e/i}$ which is given
1378: by \eqsref{isotropic} and (\ref{ions}).  [In fact, we compute the
1379: electron-electron flux by numerically evaluating the integrals in
1380: \eqsref{iso-gen}.]  To this is added
1381: the quasilinear flux from \eqsref{wave-flux} and (\ref{d-ql}) and the
1382: electric-field-induced flux from \eqref{s-e}.  Both these terms are
1383: converted into spherical coordinates using \eqsref{transform}.  The
1384: total flux is then given by the general equations (\ref{spherical}b)
1385: and (\ref{spherical}c).
1386: 
1387: The diffusion and friction coefficients are computed at the points at
1388: which we need to know $S_v$ and $S_\theta$.  Thus we compute
1389: $D_{vv,i+1/2,j}$, $D_{v\theta,i+1/2,j}$, $F_{v,i+1/2,j}$, and
1390: $D_{\theta v,i,j+1/2}$, $D_{\theta\theta,i,j+1/2}$,
1391: $F_{\theta,i,j+1/2}$.
1392: The coefficients for $S_v$ are not required at $j=0$, nor
1393: those for $S_\theta$ at $i=0$, $M$, because these fluxes are multiplied
1394: by zero in \eqref{flux-grid}.  The boundary conditions \eqsref{bc-e-spher}
1395: at $v_{\rm max}$ are handled by setting
1396: 	$$\eqalign{
1397: D_{vv,i+1/2,N}&\leftarrow0,\cr
1398: D_{v\theta,i+1/2,N}&\leftarrow0,\cr
1399: F_{v,i+1/2,N}&\leftarrow\max(F_{v,i+1/2,N},0).\cr}$$
1400: 
1401: Next we must specify the way in which $f$ and its derivatives are to be
1402: computed at the edges of the cells---i.e., locations $(i+1/2,j)$ and
1403: $(i,j+1/2)$---in terms of the values of $f$ at the centers of the cells
1404: $(i+1/2,j+1/2)$.  Two of the terms are straightforward:
1405: 	$$\eqalignno{
1406: {\partial f_{i+1/2,j}\over\partial v}&
1407: ={f_{i+1/2,j+1/2}-f_{i+1/2,j-1/2}\over \Delta v},&(\eqlab{deriv}a)\cr
1408: {\partial f_{i,j+1/2}\over\partial \theta}&
1409: ={f_{i+1/2,j+1/2}-f_{i-1/2,j+1/2}\over \Delta
1410: \theta}.&(\ref{deriv}b)\cr}$$
1411: 	Again these expressions are accurate to second order.
1412: 
1413: The evaluation of $f$ at the cell edges uses a method proposed by
1414: Chang and Cooper \cite{Chang} extended here to
1415: two dimensions.  The simple method, i.e., 
1416: $$f_{i+1/2,j}=
1417: \half(f_{i+1/2,j+1/2}+f_{i+1/2,j-1/2}),$$  turns out to give poor
1418: results for the steady-state distribution.  Chang and Cooper replace
1419: this with
1420: 	$$\eqalignno{
1421: f_{i+1/2,j}&=(1-\delta_{i+1/2,j}) f_{i+1/2,j+1/2}
1422: +\delta_{i+1/2,j} f_{i+1/2,j-1/2},&(\eqlab{value}a)\cr
1423: f_{i,j+1/2}&=(1-\delta_{i,j+1/2}) f_{i+1/2,j+1/2}
1424: +\delta_{i,j+1/2} f_{i-1/2,j+1/2},&(\ref{value}b)\cr}$$
1425: 	where the $\delta$s are given by
1426: 	$$\eqalignno{
1427: \delta_{i+1/2,j}&=g(-\Delta v F_{v,i+1/2,j}/D_{vv,i+1/2,j}),
1428: &(\eqlab{del-def}a)\cr
1429: \delta_{i,j+1/2}&=g(-\Delta \theta F_{\theta,i,j+1/2}/D_{\theta\theta,i,j+1/2}),
1430: &(\ref{del-def}b)\cr}$$
1431: and
1432: 	$$g(w)={1\over w}-{1\over \exp(w)-1}.\Eqlab{g-def}$$
1433: 
1434: The role of the $\delta$ is to weight the averaging performed in
1435: \eqsref{value}.  The weighting is needed because often $f$ is a strongly
1436: (exponentially) varying function of ${\bf v}$.  An acute example of this
1437: is the Maxwellian distribution which varies very strongly for large
1438: $v$.  In fact, the weighting is such that a Maxwellian is an {\sl exact}
1439: steady-state solution when there is no rf and no electric field and
1440: when $C_{\rm Max}^{e/e}$ is employed as the electron-electron collision
1441: operator.  This is easily seen because for any isotropic distribution
1442: $S_{c\theta}=0$; in that case, we also require $S_{cv}=0$ in the steady
1443: state (because there are no sources or sinks of electrons).  Using
1444: eqs.~(\ref{isotropic}a) (with $a=b=e$), (\ref{deriv}a), and
1445: (\ref{value}a), together with
1446: $F_{cv,i+1/2,j}^{e/e}/D_{cvv,i+1/2,j}^{e/e}=-v_j$, we find
1447: 	$${f_{i+1/2,j+1/2}\over f_{i+1/2,j-1/2}}
1448: ={f_{m,j+1/2}\over f_{m,j-1/2}}
1449: =\exp(-v_j\,\Delta v).$$
1450: 	The errors in various moments of $f$ are, therefore, exponentially
1451: small.  With one-dimensional equations the weighting cures the problem
1452: of $f$ becoming negative \cite{Chang}.  With our two-dimensional
1453: equation, this problem is alleviated but not cured.  In general, this
1454: problem is solved by taking a sufficiently fine mesh (assuming that the
1455: electron-electron collision operator preserves the non-negative nature
1456: of $f$).
1457: 
1458: The function $g$ has the properties
1459: 	$$\eqalign{g(w)&=1-g(-w),\cr
1460: g(w)&={1\over 2}-{w\over 12}+{w^3\over 720}+\ldots,\cr
1461: g(-\infty)&=1,\qquad
1462: g(0)={1\over 2},\qquad g(\infty)=0.\cr}$$
1463: 	The first two properties are useful for evaluating $g(w)$ for
1464: $w\gg1$ and $w\approx 0$, respectively.
1465: 
1466: The values of the cross-derivative terms which multiply the
1467: off-diagonal terms in the diffusion tensor ($D_{v\theta}$ and
1468: $D_{\theta v}$) are now given in terms of \eqsref{value} as
1469: 	$$\eqalignno{
1470: {\partial f_{i+1/2,j}\over \partial \theta}&=
1471: {f_{i+3/2,j}-f_{i-1/2,j}\over 2\Delta\theta},&(\eqlab{cross}a)\cr
1472: {\partial f_{i,j+1/2}\over \partial v}&=
1473: {f_{i,j+3/2}-f_{i,j-1/2}\over 2\Delta v}.&(\ref{cross}b)\cr}$$
1474: 
1475: The internal boundary conditions \eqsref{spher-bc} give the values of
1476: $f_{i+1/2,j+1/2}$ beyond the internal boundaries as
1477: 	$$\eqalign{f_{i+1/2,-1/2}&=f_{M-i-1/2,1/2},\cr
1478: f_{-1/2,j+1/2}&=f_{1/2,j+1/2},\cr
1479: f_{M+1/2,j+1/2}&=f_{M-1/2,j+1/2}.\cr}$$
1480: 	These conditions are only needed in the evaluation of
1481: cross-derivative terms.  The form of \eqref{flux-grid} automatically
1482: takes care of the internal boundaries for the other terms.
1483: 
1484: The external boundary at $v=v_{\rm max}$ is treated as follows: In the
1485: computation of $S_{v,i+1/2,N}$ we need only worry about the friction
1486: term (since $D_{vv}=D_{v\theta}=0$ on the boundary) so that only
1487: $f_{i+1/2,N}$ is needed.  Furthermore, the friction coefficient
1488: $F_{v,i+1/2,N}$ is non-negative.  From eq.~(\ref{value}a), we have
1489: $f_{i+1/2,N}=f_{i+1/2,N-1/2}$ because $\delta_{i+1/2,N}\rightarrow1$
1490: for $F_{v,i+1/2,N}>0$ and $D_{vv,i+1/2,N}=0+$.  (Obviously the value of
1491: $f_{i+1/2,N}$ is not required where $F_{v,i+1/2,N}=0$.)  Recall that the
1492: equation reduces to hyperbolic type on this boundary, so that no
1493: boundary condition should need to be specified here, as indeed is the
1494: case.  In fact, the method reduces to the standard upstream differencing
1495: for a hyperbolic equation on this boundary.  In the computation of
1496: $S_{\theta,i,N-1/2}$, only the cross-derivative term $\partial
1497: f_{i,N-1/2}/ \partial v$ potentially involves points outside the
1498: integration domain.  In this term, we use
1499: 	$${\partial f_{i,N-1/2}\over \partial v}=
1500: {f_{i,N-1/2}-f_{i,N-3/2}\over \Delta v},$$
1501: 	instead of eq.~(\ref{cross}b).
1502: 
1503: \subsection{Matrix formulation}
1504: 
1505: For collisions off a Maxwellian background the problem is {\sl linear}
1506: so that \eqref{flux-grid} can be rewritten as
1507: $${\partial f\over \partial t}+ Af=h,\Eqlab{matrix-eq}$$
1508: 	where $f$ is a vector of length $MN$ of the values
1509: $f_{i+1/2,j+1/2}$ and $A$ is an $MN\times MN$ matrix of coefficients.
1510: The right-hand side $h$ (also a vector of length $MN$) is inserted to
1511: aid in the treatment of other collision operators.  For the Maxwellian
1512: collision operator, we have $h=0$.  It is convenient to split $A$ into
1513: three pieces, namely
1514: 	$$A=A_v +A_\theta +A_\times,$$
1515: 	where $A_v$ contains the terms proportional to $D_{vv}$ and
1516: $F_v$, $A_\theta$ contains those proportional to $D_{\theta\theta}$
1517: and $F_\theta$, and $A_\times$ contains the cross-derivative terms
1518: proportional to $D_{v\theta}$ and $D_{\theta v}$.  With the difference
1519: scheme given in this section $A_v$ and $A_\theta$ are tridiagonal
1520: matrices.  Thus we can write
1521: 	$$\eqalignno{(A_vf)_{i+1/2,j+1/2}
1522: &= a_{v,i+1/2,j+1/2}f_{i+1/2,j-1/2}
1523:  +b_{v,i+1/2,j+1/2}f_{i+1/2,j+1/2}\cr
1524: &\quad +c_{v,i+1/2,j+1/2}f_{i+1/2,j+3/2},&(\eqlab{tridiag}a)\cr
1525: (A_\theta f)_{i+1/2,j+1/2}
1526: &= a_{\theta,i+1/2,j+1/2}f_{i-1/2,j+1/2}
1527:  +b_{\theta,i+1/2,j+1/2}f_{i+1/2,j+1/2}\cr
1528: &\quad +c_{\theta,i+1/2,j+1/2}f_{i+3/2,j+1/2},&(\ref{tridiag}b)\cr}$$
1529: where
1530: $$\eqalignno{
1531: a_{v,i+1/2,j+1/2}&=
1532: {v_j^2\over B_v}
1533: \biggl(-{D_{vv,i+1/2,j}\over\Delta v}
1534: -F_{v,i+1/2,j}\delta_{i+1/2,j}\biggr),&(\eqlab{mat-coeff}a)\cr
1535: b_{v,i+1/2,j+1/2}&=
1536: {v_j^2\over B_v}
1537: \biggl({D_{vv,i+1/2,j}\over\Delta v}
1538: -F_{v,i+1/2,j}\epsilon_{i+1/2,j}\biggr)\cr
1539: &\quad{}+{v_{j+1}^2\over B_v}
1540: \biggl({D_{vv,i+1/2,j+1}\over\Delta v}
1541: +F_{v,i+1/2,j+1}\delta_{i+1/2,j+1}\biggr),&(\ref{mat-coeff}b)\cr
1542: c_{v,i+1/2,j+1/2}&=
1543: {v_{j+1}^2\over B_v}
1544: \biggl(-{D_{vv,i+1/2,j+1}\over\Delta v}
1545: +F_{v,i+1/2,j+1}\epsilon_{i+1/2,j+1}\biggr),&(\ref{mat-coeff}c)\cr
1546: a_{\theta,i+1/2,j+1/2}&=
1547: {\sin\theta_i\over B_\theta}
1548: \biggl(-{D_{\theta\theta,i,j+1/2}\over v_{j+1/2}\Delta \theta}
1549: -F_{\theta,i,j+1/2}\delta_{i,j+1/2}\biggr),&(\ref{mat-coeff}d)\cr
1550: b_{\theta,i+1/2,j+1/2}&=
1551: {\sin\theta_i\over B_\theta}
1552: \biggl({D_{\theta\theta,i,j+1/2}\over v_{j+1/2}\Delta \theta}
1553: -F_{\theta,i,j+1/2}\epsilon_{i,j+1/2}\biggr)\cr
1554: &\quad{}+{\sin\theta_{i+1}\over B_\theta}
1555: \biggl({D_{\theta\theta,i+1,j+1/2}\over v_{j+1/2}\Delta \theta}
1556: +F_{\theta,i+1,j+1/2}\delta_{i,j+1/2}\biggr),&(\ref{mat-coeff}e)\cr
1557: c_{\theta,i+1/2,j+1/2}&=
1558: {\sin\theta_{i+1}\over B_\theta}
1559: \biggl(-{D_{\theta\theta,i+1,j+1/2}\over v_{j+1/2}\Delta \theta}
1560: +F_{\theta,i+1,j+1/2}\epsilon_{i,j+1/2}\biggr),&(\ref{mat-coeff}f)\cr}$$
1561: 	where $\epsilon=1-\delta$, $B_v=\Delta v\,v_{j+1/2}^2$, and
1562: $B_\theta=v_{j+1/2}\Delta \theta\,\sin\theta_{i+1/2}$.  With these
1563: coefficients the boundary conditions are reflected in the relations
1564: $a_{v,i+1/2,1/2}=c_{v,i+1/2,N-1/2}=0$ and
1565: $a_{\theta,1/2,j+1/2}=c_{\theta,N-1/2,j+1/2}=0$, which are automatically
1566: satisfied.
1567: 
1568: The matrix $A_\times$ is more complicated with $(A_\times
1569: f)_{i+1/2,j+1/2}$ depending, in general, on the eight nearest neighbors
1570: to $f_{i+1/2,j+1/2}$.  The boundary conditions have to be explicitly
1571: included in this matrix.  We do not give expressions for the components
1572: of $A_\times$ here because only the product $A_\times f$ is ever needed
1573: in the calculation.  This is most easily computed directly in terms
1574: of the flux; this also cuts down on the storage requirements.
1575: 
1576: \subsection{Alternate collision operators}
1577: 
1578: The methods we will describe in the next sections for solving
1579: \eqref{matrix-eq} depend on the linearity of this equation and 
1580: the fact that $A_v$ and $A_\theta$ are tridiagonal matrices.  With
1581: more complicated electron-electron collision operators, these
1582: conditions no longer hold.  However, the techniques can still be used
1583: because the difference between the other collision terms and the
1584: Maxwellian collision term varies slowly in time.
1585: 
1586: If the full electron-electron collision operator is used, the basic
1587: framework given above still applies, except that the diffusion and
1588: friction coefficients $\mat D_c^{e/e}$ and ${\bf F}_c^{e/e}$ are now
1589: given in terms of gradients of the Rosenbluth potentials
1590: \eqsref{coll-ros}.  These coefficients depend on $f$ making the equation
1591: nonlinear.  In practice, the dependence on $f$ is weak so that the
1592: coefficients only need to be recomputed occasionally.  This also means
1593: that the equation is approximately linear so that the linear matrix
1594: techniques used to advance the equation in time still apply.
1595: 
1596: If the linearized or truncated collision operators are used, then the
1597: equation remains linear but with a term which involves an integral over
1598: $f$, namely $C\bigl(f_m(v),f({\bf v})\bigr)$ or the truncation of this term.
1599: Again, this term is weakly dependent on $f$ so that it need not be
1600: recomputed every time step.  It is then most convenient to regard this
1601: term as the inhomogeneous driving term $h$ \eqref{matrix-eq}.  For
1602: the truncated collision operator $C_{\rm trunc}^{e/e}$,
1603: \eqref{truncated}, the elements of $h$ are given
1604: by $C\bigl(f_m(v),f^{(1)}(v)\cos\theta\bigr)$ evaluated at
1605: $(v_{j+1/2},\theta_{i+1/2})$.  The computation of this term is described
1606: in \appref{numerical}.
1607: 
1608: \section{Time Differencing}\label{time}
1609: 
1610: \subsection{Crank--Nicholson method}
1611: 
1612: We now turn to the method for advancing the Fokker--Planck equation in
1613: time.    If the time step is $\Delta t$, then we define
1614: 	$$f^k=f(t=t_k),\qquad t_k=k\Delta t.\Eqlab{time-step}$$
1615: 	The simplest way of advancing \eqref{matrix-eq} is the explicit
1616: scheme
1617: 	$${f^{k+1}-f^k\over \Delta t}+Af^k=h.$$
1618: 	This is only accurate to first order in $\Delta t$.  Furthermore,
1619: $\Delta t$ must be chosen to be very small, on the order of $\Delta v^2$ or
1620: $\Delta \theta^2$, for stability.  These defects are easily remedied by
1621: the Crank--Nicholson scheme \cite{Marchuk} which reads
1622: 	$${f^{k+1}-f^k\over \Delta t}+A{f^{k+1}+f^k\over 2}=h.
1623: \Eqlab{crank}$$
1624: 	This scheme is accurate to second order in $\Delta t$ and
1625: is stable if $A$ is positive definite.  (This is a
1626: condition possessed by the continuous form of the operator $A$.)  In
1627: order to solve \eqref{crank} for $f^{k+1}$ we have to compute the
1628: inverse of $(I+\half\Delta t\,A)$.  This is a large banded matrix which
1629: can either be inverted using iterative methods or using Gaussian
1630: elimination.  In both cases the number of operations is $O(N^3)$,
1631: (assuming $M\sim N$) making it a very expensive proposition.
1632: (This approach is discussed further in \secref{steady-state}.)
1633: 
1634: \subsection{Alternating-direction-implicit method}
1635: 
1636: Although $(I+\half\Delta t\,A)$ is difficult to invert, the matrices
1637: $(I+\half\Delta t\,A_v)$ and $(I+\half\Delta t\,A_\theta)$ are rather
1638: easily inverted.  This allows the {\sl alternating-direction-implicit}
1639: method \cite{Marchuk} to be used.  Unfortunately, $(I+\half\Delta
1640: t\,A_\times)$ is not easily inverted and this means that the
1641: cross-derivative terms are treated explicitly in this method.  Consider
1642: the equation
1643: 	$$\biggl(I+{\Delta t\over2}A_v\biggr)
1644: \biggl(I+{\Delta t\over2}A_\theta\biggr){f^{k+1}-f^k\over\Delta t}
1645: +Af^k=h.\Eqlab{split}$$
1646: 	If we rearrange the terms in this equation to give
1647: 	$$\biggl(I+{\Delta t^2\over 4}A_vA_\theta\biggr)
1648: {f^{k+1}-f^k\over\Delta t}
1649: +(A_v+A_\theta){f^{k+1}+f^k\over 2}+A_\times f^k=h,$$
1650: 	we see that this method differs from the Crank--Nicholson
1651: method in two respects.  Firstly, there is a $\Delta t^2$ term
1652: multiplying the time difference term.  This difference is unimportant
1653: because it does not alter the accuracy of the scheme.  Secondly, the
1654: cross-derivative terms are treated explicitly.  If we ignore the
1655: cross-derivative terms, \eqref{split} is as accurate as the
1656: Crank--Nicholson scheme, but is much easier to realize because it is
1657: easy to solve \eqref{split} for $f^{k+1}$.  The explicit treatment of the
1658: cross-derivative terms lowers the accuracy and the stability, putting a
1659: limit on the maximum $\Delta t$ that can be used.  On the other hand,
1660: the implicit treatment of the other terms means that this method is far
1661: superior to the fully explicit method.
1662: 
1663: We can compute $f^{k+1}$ from \eqref{split} in a series of simple steps:
1664: $$\eqalign{\phi^k&=h-Af^k,\cr
1665: \xi^{k+1/2}&=\biggl(I+{\Delta t\over2}A_v\biggr)^{-1}\phi^k,\cr
1666: \xi^{k+1}&=\biggl(I+{\Delta t\over2}A_\theta\biggr)^{-1}\xi^{k+1/2},\cr
1667: f^{k+1}&=f^k+\Delta t\,\xi^{k+1}.\cr}$$
1668: 	The inversion of the matrices is carried out using Gaussian
1669: elimination as described in \appref{numerical}.
1670: 
1671: \subsection{Example}
1672: 
1673: Let us consider a specific example relevant to lower hybrid current
1674: drive.  The plasma consists of electrons and infinitely massive ions
1675: with $Z_i=1$.  Electron-electron collisions are computed assuming a
1676: Maxwellian background using $C_{\rm Max}^{e/e}$ \eqref{coll-max}.
1677: Electron-ion collisions are given by \eqref{coll-ions}.  The effect of
1678: the lower hybrid waves is modeled by a quasilinear diffusion
1679: coefficient given by \eqsref{d-ql} with $D_0=1$, $v_1=3$, and $v_2=5$.
1680: The electric field $E$ is taken to be zero.  Except for minor details
1681: this is the same example treated in the paper on lower hybrid
1682: current drive \cite{Karney-lh}.  (The time normalization used
1683: in that paper differs from the one adopted here by a factor of two.)  We
1684: take $f(t=0)=f_m$, $v_{\rm max}=10$, $M=N=100$, and $\Delta t=0.2$.
1685: 
1686: In studies of current drive, we are principally interested in the
1687: current density $J$, the rf power absorbed per unit volume by the
1688: plasma $P$, and their ratio $J/P$.  These are defined by
1689: \eqsref{curr-def} and (\ref{pd-def}) whose discrete forms read
1690: 	$$\eqalignno{
1691: J&={\numint(v\cos\theta)\over n},&(\eqlab{curr-def-grid})\cr
1692: P&={1\over n}\sum_{i=0}^{M-1}\sum_{j=0}^{N}
1693: 2\pi\sin\theta_{i+1/2} v_j^3 S_{wv,i+1/2,j} \,\Delta v\Delta\theta,
1694: &(\eqlab{pd-def-grid})\cr}$$
1695: 	    where $n$ is given by \eqref{n-grid}.  (These definitions
1696: include a $1/n$ factor, because the $n$ is included in the
1697: normalizations for $J$ and $P$.)
1698: 
1699: The current is plotted as a function of time in
1700: \figref{current}.  With $\Delta t=0.5$, the integration is unstable.  The
1701: difference in the values of the current when the equations are
1702: integrated with $\Delta t=0.2$ and $\Delta t=0.05$ is about $0.1\%$ of
1703: the final current.
1704: 
1705: The steady-state solution for $f$ is shown in \figref{f-steady}.  This
1706: may be obtained by integrating the equation sufficiently long (until
1707: about $t=1000$) with a fixed time step or else using the techniques
1708: described in \secref{steady-state}.  (With this numerical method, the
1709: steady state is independent of $\Delta t$.)  The plateau in the resonant
1710: region is clearly visible as well as the considerable perpendicular
1711: heating.  Using \eqsref{curr-def-grid} and (\ref{pd-def-grid}), we have
1712: $J=5.754\times10^{-2}$, $P=4.011\times10^{-3}$, and $J/P=14.34$.
1713: 
1714: The flux plot for this case is given in \figref{flux-fig}.  This shows
1715: that the combination of rf diffusion and collisional scattering induces
1716: a perpendicular flux in the resonant region.  Such flux plots are
1717: useful in providing guidance for the analytic solution of this problem
1718: \cite{asymp}.
1719: More extensive examination of this example can be found in the original
1720: paper \cite{Karney-lh} including projections onto the $v_\parallel$
1721: axis, slices at constant $v_\perp$, etc.
1722: 
1723: There are two possible sources of error in these results: errors
1724: arising from the finite boundary (i.e., because $v_{\rm max}$ is
1725: finite) and errors arising from the finite mesh.  The effect of the
1726: boundary can be determined by increasing $v_{\rm max}$ to 20 (and
1727: increasing $N$ to 200).  In the steady state, this gives
1728: $J=5.759\times10^{-2}$, $P=4.012\times10^{-3}$, $J/P=14.35$---changes
1729: of less than $0.1\%$.  Thus for this particular problem, $v_{\rm
1730: max}=10$ is adequate.
1731: 
1732: The effect of the discrete spatial grid is found by varying $\Delta v$
1733: and $\Delta \theta$.  This we do by keeping $v_{\rm max}=10$ varying $M$
1734: and $N$ with $M=N$.  Thus we have $N=10/\Delta v$ and
1735: $\Delta\theta=\Delta v \,\pi/10$.  The results for $J$ and $J/P$ are
1736: shown in \figref{conv-d}.  We see that there is a lot of scatter in the
1737: data which arises because $\mat D_w$ is discontinuous.  As $\Delta v$
1738: and $\Delta \theta$ are varied, grid points (those on which the flux is
1739: defined) enter or leave the resonant region $v_1<v_\parallel<v_2$.  Each
1740: time this happens, there is a jump in $J$ and $P$.  As $\Delta
1741: v\rightarrow 0$, $J$ approaches its asymptotic value of about
1742: $5.6\times10^{-2}$ and the convergence to this value is as $\Delta v$.
1743: The finite mesh error in $J$ with $M=N=100$ is about $3\%$.  This rate
1744: of convergence can be understood because $J$ and $P$ are exponentially
1745: dependent on $v_1$ [$J\sim \exp(-\half v_1^2)$] and $v_1$ is determined
1746: only to within $\pm\half\Delta v$.  Thus the relative error in $J$ and
1747: $P$ is about $\exp(\half v_1\Delta v)-1\approx \half v_1\Delta v$.  This
1748: gives a relative error of $15\%$ for $v_1=3$, $N=100$, $v_{\rm
1749: max}=10$.  The actual error is somewhat less than this because the
1750: boundary of the resonant region cuts across the grid lines and so $v_1$
1751: is in fact determined more accurately than was assumed here.  Because
1752: $J$ and $P$ are both subject to the same error, the ratio $J/P$ is more
1753: accurately given: convergence to the asymptotic value of $14.24$ is as
1754: $\Delta v^2$ and the value with $M=N=100$ is in error by less than $1\%$.
1755: 
1756: If instead we use the truncated electron-electron collision operator $C_{\rm
1757: trunc}^{e/e}$, the steady-state distribution function is
1758: rather similar to that shown in \figref{f-steady}.  However, the flux
1759: plot \figref{trunc-flux} shows a new eddy at low velocities due to the
1760: overall drift of the electrons with respect to the ions.  (This plot is
1761: obtained with the same parameters as for \figref{flux-fig}.)  In this
1762: case, we find $J=7.092\times10^{-2}$, $P=4.294\times10^{-3}$,
1763: $J/P=16.52$.  The enhancement of the efficiency $J/P$ comes about
1764: because momentum (and hence current) is no longer lost when tail
1765: electrons collide with bulk electrons.
1766: 
1767: A check on the implementation of the $C_{\rm trunc}^{e/e}$ is given by
1768: measuring the electrical conductivity.  For $Z_i=1$, the exact
1769: conductivity is given by \tabref{conduct} as $J/E=7.429\approx
1770: 0.582\times16\sqrt{2/\pi}$ \cite{Spitzer}.  Integrating the
1771: Fokker--Planck equation using the truncated collision operator with no
1772: rf $D_0=0$ and a small electric field $E=10^{-3}$, the conductivity is
1773: $J/E=7.446$, a $0.3\%$ error.  This small error is probably attributable
1774: partly to the finite mesh size (here we again took $M=N=100$ and
1775: $v_{\rm max}=10$) and partly to the finiteness of $E$ (since there is a
1776: contribution to the current which varies as $E^3$).  In contrast, if
1777: $C_{\rm Max}^{e/e}$, is used the conductivity is $J/E=3.772$ a factor of
1778: two too small \cite{Kulsrud}.
1779: 
1780: \section{Steady-State Solution}\label{steady-state}
1781: 
1782: \subsection{Statement of problem}
1783: 
1784: Often, we are only interested in the steady-state solution to
1785: the Fokker--Planck equation.  Nearly always we must resort to an
1786: iterative method for obtaining the steady state.  In that case we need
1787: some measure of how close we are to the steady state so that iteration
1788: may be stopped when this is small enough.  The measure we shall employ is
1789: 	$$R={1\over n}
1790: \sqrt{\numint\biggl[\biggl({\partial f\over\partial t}\biggr)^2\biggr]},
1791: \Eqlab{residue}$$
1792: 	where the {\sl residue} $\partial f/\partial t$ is given by
1793: \eqref{flux-grid}.  Somewhat arbitrarily we use $R=10^{-9}$ as the
1794: convergence criterion.
1795: 
1796: One obvious way of obtaining a steady state is to integrate the
1797: time-depen\-dent solution as described in \secref{time} for a long time.
1798: This should be done with the largest time step consistent with
1799: stability.  For the example shown in \figref{f-steady}, the
1800: convergence criterion is met at time $t=812$.  The largest time step
1801: that can be used is approximately $0.2$; so that 4060 steps are
1802: required.  The {\sc CPU} time required to run the Fokker--Planck code on
1803: the Cray--1 is approximately $2\,\mu \rm s$ per mesh point per time
1804: step.  Thus, achieving the steady state by this method takes about
1805: $80\,\rm s$.  This is rather expensive and it is therefore desirable to
1806: find faster methods.
1807: 
1808: However, this method is quite effective when
1809: $A_\times=0$.  Then the numerical scheme is stable even if $\Delta t$ is
1810: large.  For example, for the electric field example discussed in
1811: \secref{time} in which $D_0=0$ and $E=10^{-3}$, we can take $\Delta t
1812: = 1$, and the convergence criterion is met after 220 steps.  Here
1813: the integral portion of $C_{\rm trunc}^{e/e}$, which is represented by
1814: the term $h$ in \eqref{matrix-eq}, is evaluated every tenth time step.
1815: The numerical method is stable for larger values of $\Delta t$.  But,
1816: because the integration is less accurate, more steps are required to
1817: meet the convergence criterion.  With large $\Delta t$ the numerical
1818: solution tends to oscillate about the steady state.
1819: 
1820: \subsection{Chebyshev acceleration}
1821: 
1822: A significant improvement can be achieved by using a varying time step.
1823: Hewett {\sl et al.}\ \cite{Hewett} describe an adaptive time selection
1824: for the alternating direction implicit method which speeds the
1825: convergence by a factor of two to three.  Here we describe Chebyshev
1826: acceleration \cite{Marchuk} which is a nonadaptive method for
1827: selecting varying time steps.  We choose the time step
1828: $\Delta t_k=t_{k+1}-t_k$ according to 
1829: 	$$\Delta t_k={2\over
1830: \displaystyle
1831: \beta+\alpha-(\beta-\alpha)\cos\biggl({[2(k\bmod K)+1]\pi
1832: \over 2K}\biggr)},
1833: \Eqlab{Tcheb-dt}$$
1834: 	where $\alpha$, $\beta$, and $K$ are constants with
1835: $\alpha<\beta$ and $K={\rm integer}$.  The advantage of this method is
1836: that by changing a few lines of code it can easily be incorporated into
1837: the alternating-direction-implicit method described in \secref{time}.
1838: A fixed time step is recovered in the special case
1839: $\alpha=\beta=1/\Delta t$.
1840: 
1841: Let us discuss the choice of the parameters in \eqref{Tcheb-dt}.  With
1842: $K$ large, \eqref{Tcheb-dt} gives a series of $K$ time steps (repeated
1843: periodically) varying from $1/\alpha$ down to $1/\beta$.  In the
1844: examples we consider, we take $K=20$.  Then the maximum time
1845: step is somewhat less than $1/\alpha$ while the minimum time step is
1846: very close to $1/\beta$.  In order to realize performance gains with
1847: this method we wish to pick the minimum time step comfortably within
1848: the stability threshold for the fixed-time-step method, while the
1849: maximum time step is considerably greater than the stability
1850: threshold.
1851: 
1852: The method works because the long wavelength eigenmodes of the linear
1853: operator decay slowly but are stable with large $\Delta t$; on the
1854: other hand, the
1855: short wavelength modes decay rapidly but are only stable if $\Delta t$
1856: is small.  Consider a particular cycle of $K$ steps.  During the initial
1857: large time steps, the long wavelength modes are efficiently damped
1858: (because $\Delta t$ is large), but the short wavelength modes grow.
1859: This is followed by successively shorter time steps which damp the
1860: short wavelength modes.
1861: 
1862: For the example shown in \figref{f-steady}, the stability threshold for
1863: $\Delta t$ lies between 0.2 and 0.5.  Thus we choose $1/\beta=0.05$ and
1864: $1/\alpha=1000$.  With $K=20$ this gives a maximum time step of 31.4, a
1865: minimum step of 0.05, and an average time step of 1.95.  Since the
1866: average time step is about 10 times the largest time step that can be
1867: used in the fixed time step scheme, we expect convergence to be 10
1868: times faster.  Indeed this is the case.  The convergence criterion is met
1869: after 400 steps at $t=790$.  This takes about $8\,\rm s$ of {\sc CPU}
1870: time.  The variation of $R$ with time is shown in \figref{residu-fig}.
1871: This shows the growth of $R$ during the large time steps followed by a
1872: drop in $R$ as the instabilities are quenched during the small time
1873: steps.  The overall decay of $R$ with $t$ closely matches that seen with
1874: a fixed time step.  (This is contrary to the experience of Hewett {\sl
1875: et al.}\ with their adaptive code in which the rates of decay are very
1876: different \cite{Hewett}.)
1877: 
1878: \subsection{Runaway problem}
1879: 
1880: If the electric field is sufficiently large to produce runaways, i.e.,
1881: $E>v_{\rm max}^{-2}$, then as $t\rightarrow\infty$ a steady state is
1882: reached which decays at the runaway rate $\gamma$ (assuming that a
1883: linear collision operator is employed).  Because $f$ and all its moments
1884: decay at the same rate, $\gamma$ is given from \eqref{cons-grid} as
1885: 	$$\gamma={\flux 1\over \numint 1},\Eqlab{gamma-def}$$
1886: which we will take to be the definition of $\gamma$ for all $t$.
1887: Thus we write 
1888: 	$$f({\rm v},t)=f'({\rm v},t)
1889: \exp\biggl(-\int_0^t\gamma(t')\, dt'\biggr),\Eqlab{f-decay}$$
1890: 	where $\gamma$ is given by \eqref{gamma-def} and
1891: $f'(t\rightarrow\infty)$ is independent of $t$.
1892: 	If \eqref{f-decay} is substituted into \eqref{matrix-eq}, we
1893: obtain
1894: 	$${\partial f'\over \partial t}+ (A-\gamma)f'=0,\Eqlab{matrix-mod}$$
1895: 	where for simplicity we set the inhomogeneous term $h$ to zero.
1896: Because $\gamma$ is expressed as an integral over $f$
1897: \eqref{gamma-def}, it varies slowly and need not be evaluated very
1898: often.  Thus \eqref{matrix-mod} may be regarded as a linear equation and
1899: solved in precisely the same way as \eqref{matrix-eq} (with $h=0$)
1900: except that $\gamma$ must be subtracted from $b_{v,i+1/2,j+1/2}$
1901: eq.~(\ref{mat-coeff}b).
1902: 
1903: As an example, \figref{run-fig} shows the steady-state distribution
1904: obtained by this method with $Z_i=1$,
1905: $E=0.06$, $M=N=100$, $v_{\rm max}=10$, and
1906: electron-electron collisions given by $C_{\rm Max}^{e/e}$.  Since there
1907: is no rf diffusion term, there are no cross-derivative terms and the
1908: steady state is most easily obtained by taking a constant time step of
1909: $\Delta t=1$.  The runaway rate $\gamma$ is recomputed every ten time
1910: steps and the convergence condition $R=10^{-9}$ is met after 820 time
1911: steps.  In the steady state, we have $\gamma=5.211\times10^{-5}$ and
1912: $J=0.3133$.  These are close to the results obtained by Kulsrud {\sl et
1913: al.}\ \cite{Kulsrud}, namely $\gamma=5.411\times10^{-5}$ and
1914: $J=0.3143$.
1915: 
1916: Again, it is important to explore the possible errors in these figures.
1917: Extending the boundary to $v_{\rm max}=20$ and doubling $N$ to $200$
1918: gives $\gamma=5.210\times10^{-5}$ and $J=0.4514$.  While there is
1919: practically no change in $\gamma$, $J$ is about $50\%$ larger.  This
1920: discrepancy arises because there is a large contribution to the total
1921: current by the runaways in the region $10<v<20$.  We can verify this by
1922: estimating the total current for an arbitrary $v_{\rm max}$ on the
1923: basis of the results from $v_{\rm max}=10$.  For simplicity, assume that
1924: all the runaways are concentrated near $v_\perp=0$.  From small $\gamma$
1925: and in the limit $t\rightarrow\infty$, the runaway distribution is
1926: independent of $v_\parallel$, so that $f(v_\parallel\gg v_t)\approx
1927: (\gamma/E) \delta({\bf v}_\perp)$.  The current obtained by integrating
1928: $v_\parallel f$ out to $v=v_{\rm max}$ is then
1929: 	$$J(v_{\rm max})\approx J_{\rm bulk}
1930: +\half(\gamma/E)v_{\rm max}^2,$$
1931: 	where, using the data from $v_{\rm max}=10$, we have $J_{\rm
1932: bulk}=0.270$.  We can interpret $J_{\rm bulk}$ as the current carried by
1933: the bulk electrons and the other term as the current carried by the
1934: runaways.  This now gives $J(v_{\rm max}=20)=0.444$ which is within
1935: $2\%$ of the observed value.  The lesson from this exercise is that it
1936: makes little sense to quote the result for $J$ when the runaway rate is
1937: appreciable because it depends strongly on $v_{\rm max}$.  It is
1938: preferable to determine the bulk current since this is then weakly
1939: dependent on $v_{\rm max}$ and has a physical interpretation.  We have
1940: seen that $v_{\rm max}=10$ is sufficiently large to give $\gamma$ and
1941: $J_{\rm bulk}$ accurately.
1942: 
1943: In order to determine the effect of the finite mesh on the runaway
1944: results, we vary $M$ and $N$ with $M=N$ and $v_{\rm max}=10$.  The
1945: results for $\gamma$ and $J$ are shown in \figref{conv-e}.  The
1946: asymptotic values are $\gamma=5.185\times10^{-5}$ and $J=0.31334$.  The
1947: errors in the values for $M=N=100$ are $0.5\%$ and $0.02\%$,
1948: respectively.  The errors are considerably less than with the rf
1949: problem in \figref{conv-d} and the convergence is much more
1950: regular (as $\Delta v^2$).
1951: 
1952: A disadvantage of solving for the decaying steady state of the
1953: distribution, \eqref{matrix-mod}, is that ${\bf S}$ is no longer
1954: divergence free.  This means that the stream lines cannot be plotted as
1955: contours of a stream function $A$, \eqref{stream-grid}.  This can be
1956: remedied by injecting electrons at the origin to match the runaway loss
1957: of particles.  Although this is a rather artificial problem, there is
1958: little error in the runaway rate provided that the runaway rate itself
1959: is small.  We implement this procedure as follows:  The loss of
1960: particles at $v=v_{\rm max}$ is
1961: 	$$n\gamma = \flux 1.$$
1962: 	We match this loss by a uniform radial flux at the origin
1963: 	$$v_0^2 S_{v,i+1/2,0} = {n\gamma \sin(\half\Delta\theta)
1964: \over 2\pi\,\Delta\theta},$$
1965: 	which is chosen to give
1966: 	$$\sum_{i=0}^{M-1}
1967: 2\pi\sin\theta_{i+1/2} v_0^2 S_{v,i+1/2,0}\,\Delta\theta = n\gamma.$$
1968: 	(The product $v_0^2 S_{v,i+1/2,0}$ is finite even though
1969: $S_{v,i+1/2,0}$ is infinite.)  From \eqref{flux-grid}, we see that this
1970: introduces a source term
1971: 	$v_0^2 S_{v,i+1/2,0}/(v_{1/2}^2\,\Delta v)$ 
1972: 	into the expressions for $\partial f_{i+1/2,1/2}/\partial t$.
1973: This is included as part of the inhomogeneous term $h$ in
1974: \eqref{matrix-eq}.  The expressions for the stream function
1975: \eqsref{stream-grid} require a slight modification to give
1976: 	$$\eqalign{A_{i,j}&=-\gamma +{2\pi v_j^2\over n}\sum_{i'=0}^{i-1}
1977: \sin\theta_{i'+1/2}S_{v,i'+1/2,j}\,\Delta\theta,
1978: \cr
1979: &=A_{i,0}-{2\pi\sin\theta_i\over n}\sum_{j'=0}^{j-1}
1980: v_{j'+1/2}S_{\theta,i,j'+1/2}\,\Delta v,
1981: \cr}$$
1982: 	where the integration constant has been chosen to given
1983: $A_{0,j}=-\gamma$ and $A_{M,j}=0$.
1984: 
1985: The flux plot computed by this method for the case shown in
1986: \figref{run-fig}, i.e., for $Z_i=1$, $E=0.06$, $M=N=100$, $v_{\rm
1987: max}=10$, is shown in \figref{run-flux}.  When computed in this way, the
1988: runaway rate is slightly lower $\gamma=5.148\times10^{-5}$ because a
1989: typical runaway particle has to be accelerated from $v=0$ instead of
1990: $v=1$.  The current $J=0.3127$ is also lower.
1991: 
1992: \subsection{Other methods}
1993: 
1994: An {\sl infinite} time step can be used if the Crank--Nicholson scheme,
1995: eq.\ (\ref{crank}), is modified so that $f^{k+1}$ is used in place of
1996: $\half(f^{k+1}+f^k)$.  Then, the steady state can be achieved in a
1997: single time step.  Of course, this entails inverting the large matrix $A$
1998: (which is why we advocated using the alternating-direction-implicit
1999: method in preference to the Crank--Nicholson method).  However, routines
2000: are available to perform such an inversion and they have been employed
2001: by O'Brien {\sl et al.} \cite{O'Brien}.  An important feature of this
2002: method is the use of disk files to hold intermediate results.
2003: (Typically, the full matrix cannot fit into memory.)  They report a {\sc
2004: CPU} time of 35 s to invert the matrix arising from the discretization
2005: of the Fokker--Planck equation on a $300\times100$ grid with this time
2006: scaling as $MN\times\min(M,N)$.  This method is therefore comparable (as
2007: far as {\sc CPU} time goes) to the Chebyshev acceleration method.  There
2008: are two potential drawbacks of this scheme: Firstly, there is a
2009: significant cost in {\sc I/O} time with this method because of the use
2010: of disk files for storage.  Secondly, the advantage of the method is
2011: reduced if the steady state cannot be reached in a single time step.
2012: This is the case with the more complicated collision operators, because
2013: the matrix $A$ is a function of time.
2014: 
2015: Various iterative methods are available for obtaining a steady-state
2016: solution \cite{Marchuk}.  These are basically approximate methods of
2017: inverting the matrix $A$.  Notable is Gauss--Seidel relaxation in which
2018: the elements of $f$ are successively updated to achieve $\partial
2019: f/\partial t=0$ at the point in question.  In line relaxation, a whole
2020: line of elements (for example, $j=\rm const$) is updated simultaneously
2021: (requiring the solution of a tridiagonal system of equations).  Line
2022: relaxation gives the same convergence rate as Gauss--Seidel relaxation
2023: and may be vectorized if the even-numbered rows ($j=\rm even$) are
2024: updated in one sweep followed by the update on the odd-numbered rows.
2025: 
2026: The odd-even line relaxation method is extended with the
2027: successive-over-relaxation method where the over-relaxation parameter
2028: $\omega$ determines how much overshoot there is beyond the value of $f$
2029: which gives $\partial f/\partial t=0$.  Unfortunately, these methods
2030: give results which are roughly the same as using fixed time
2031: steps.  For the example shown in \figref{f-steady}, with the
2032: over-relaxation parameter set to $\omega=1.4$, the convergence
2033: criterion is met after 5980 steps. (Compare this to the 4060 steps
2034: required in the fixed-time-step method.  However, one relaxation step
2035: tends to be computationally less expensive than one step of the
2036: alternating-direction-implicit method.)  For this example, the method
2037: becomes unstable with $\omega\ge1.5$.
2038: 
2039: Although by themselves relaxation methods are not very useful for this
2040: problem, they are an important ingredient in the multigrid method
2041: \cite{Brandt,multigrid}.  In this method, the problem is solved at
2042: several different grid spacings (usually differing from each other by a
2043: factor of two).  A few relaxation sweeps are carried out on the finest
2044: grid.  Because relaxation is a local method, this is very effective at
2045: damping the short wavelength modes (with wavelength comparable to grid
2046: spacing).  If relaxation is continued on the finest grid, convergence
2047: would become slower because longer wavelength modes would dominate the
2048: residue.  However, in the multigrid method, the residue is transferred
2049: onto the next coarsest grid where relaxation methods are again
2050: efficient.  This process continues recursively up to very coarse grids
2051: where either relaxation methods or direct solution methods can be used.
2052: 
2053: This method has not been implemented for the Fokker--Planck equation.
2054: However, we can estimate the time required to obtain a steady state.
2055: Each relaxation step on the finest grid gives a reduction in $R$ by
2056: about a factor of two.  (The total work at the coarser grids is at most
2057: a multiple of the work on the finest grid.)  In contrast, the mean
2058: reduction in $R$ with the Chebyshev method is by $4\%$ per step (see
2059: \figref{residu-fig}).  Thus the multigrid method will require about
2060: $\log(0.5)/\log(0.96)\approx 16$ times fewer steps---an
2061: order-of-magnitude improvement over the Chebyshev method.
2062: 
2063: \section{Relativistic Treatment}\label{relativity}
2064: 
2065: \subsection{The Fokker--Planck equation}
2066: 
2067: Fokker--Planck methods have been used to study current drive by
2068: lower hybrid waves.  In a fusion plasma, these waves will interact with
2069: electrons that travel at close to the speed of light.  In such cases, it
2070: is necessary to reformulate the equation to include relativistic
2071: effects.  The first change is that the electron distribution function
2072: is expressed in momentum rather than velocity space so that
2073: \eqref{fp-1} becomes
2074: 	$${\partial f_e\over\partial t}-\sum_s C(f_e,f_s) +
2075: \nabla\cdot{\bf S}_w+q_e{\bf E}\cdot\nabla f_e = 0, \Eqlab{fp-rel}$$
2076: 	where now the $\nabla\equiv\partial/\partial{\bf p}$ operator
2077: operates in {\sl momentum} space, ${\bf S}_w$ is the rf-induced flux in
2078: momentum space, and $f_e$ is normalized so that
2079: 	$$\int f_e({\bf p})\,\dvp=n_e.$$
2080: 	In spherical coordinates we have
2081: 	$$\nabla\cdot{\bf S}
2082: ={1\over p^2}{\partial\over\partial p}p^2S_p+
2083: {1\over p\sin\theta}{\partial\over\partial \theta}\sin\theta S_\theta,$$
2084: 	where $\cos\theta=p_\parallel/p$.
2085: 
2086: In addition, the forms of the collision term and the quasilinear
2087: diffusion term are altered.
2088: 
2089: \subsection{The relativistic collision operator}
2090: 
2091: The relativistic collision operator is given by Beliaev and Budker
2092: \cite{Beliaev}.  It can again be written as the divergence of a flux
2093: $C(f_a,f_b)=-\nabla\cdot{\bf S}_c^{a/b}$, where now we have 
2094: 	$${\bf S}_c^{a/b}={q_a^2 q_b^2 \over 8\pi\epsilon_0^2}
2095: \ln \Lambda^{a/b}\int \mat U({\bf u})\cdot
2096: \biggl(
2097: f_a({\bf p}) {\partial f_b({\bf p}')\over\partial {\bf p}'}-
2098: f_b({\bf p}') {\partial f_a({\bf p})\over\partial {\bf p}}
2099: \biggr) \,\dvp'.\Eqlab{coll-rel}$$
2100: 	The expression for $\mat U$ is rather complicated \cite{Beliaev}.
2101: However, if either the test or the background species is weakly
2102: relativistic ($p\ll m_a c$ or $p'\ll m_b c$), then $\mat U$ may be
2103: approximated by its nonrelativistic form
2104: 	$$
2105: \mat U({\bf u})={u^2\mat I - {\bf uu}\over u^3},\qquad
2106: {\bf u}={\bf v}_a-{\bf v}'_b,$$
2107: 	where ${\bf v}_s={\bf p}/m_s\gamma_s$ is the velocity of
2108: species $s$, $\gamma_s=(1+p^2/m_s^2c^2)^{1/2}$ is the relativistic
2109: correction factor, and $m_s$ is the rest mass.
2110: 
2111: Despite the resemblance of \eqref{coll-rel} to \eqref{coll-land}, this
2112: collision operator cannot be readily expressed in terms of Rosenbluth
2113: potentials.  However, considerable progress can still be made by working
2114: directly with \eqref{coll-rel}.  We restrict our attention to
2115: electron-ion and electron-electron collisions.
2116: 
2117: For collisions off infinitely massive ions, we can take the
2118: ions to be stationary $v_i'\rightarrow0$ and evaluate the integrals to
2119: give
2120: 	$$C^{e/i}\bigl(f_e({\bf p})\bigr)=
2121: \Gamma^{e/e}{Z_i\over 2v_ep^2}{1\over\sin\theta}{\partial\over\partial\theta}
2122: \sin\theta{\partial\over\partial\theta}f_e({\bf v}),\Eqlab{coll-rel-ions}$$
2123: where
2124: $$\Gamma^{a/b}={n_b q_a^2 q_b^2 \ln \Lambda^{a/b}\over 4\pi
2125: \epsilon_0^2}$$
2126: (this differs by a factor of $m_a^2$ from the definition given in
2127: \secref{prelim}).
2128: 
2129: For electron-electron collisions we start with the case of an isotropic
2130: background $C\bigl(f_e({\bf p}),\penalty0 f_e^{(0)}(p)\bigr)$.  The fluxes for
2131: this term are \cite{relpap}
2132: 	$$\eqalignno{S_{cp}^{e/e}&=
2133: -D_{cp p}^{e/e} {\partial f_e\over\partial p}
2134: +F_{cp}^{e/e}f_e,&(\eqlab{isotropic-rel}a)\cr
2135: S_{c\theta}^{e/e}&=
2136: -D_{c\theta \theta}^{e/e}{1\over p}{\partial f_e\over\partial \theta}
2137: ,&(\ref{isotropic-rel}b)\cr}$$
2138: 	where
2139: 	$$\eqalignno{
2140: D_{cpp}^{e/e}&={4\pi \Gamma^{e/e}\over 3n_e}
2141: \biggl(\int_0^p p'{}^2f_e^{(0)}(p') {v_e'{}^2\over v_e^3}\,dp'+
2142: \int_p^\infty p'{}^2f_e^{(0)}(p') {1\over v_e'}\,dp'\biggr),
2143: &(\eqlab{iso-rel-gen}a)\cr
2144: D_{c\theta\theta}^{e/e}&={4\pi \Gamma^{e/e}\over 3n_e}
2145: \biggl(\int_0^p p'{}^2f_e^{(0)}(p') {3v_e^2-v_e'{}^2\over 2v_e^3}
2146: \,dp'+
2147: \int_p^\infty p'{}^2f_e^{(0)}(p') {1\over v_e'}\,dp'\biggr),
2148: &(\ref{iso-rel-gen}b)\cr
2149: F_{cp}^{e/e}&=-{4\pi \Gamma^{e/e}\over 3n_e}
2150: \biggl(\int_0^p p'f_e^{(0)}(p') {3v_e'-v_e'{}^3/c^2\over v_e^2}
2151: \,dp'+
2152: \int_p^\infty p'f_e^{(0)}(p') 2v_e/c^2\,dp'\biggr).\qquad
2153: &(\ref{iso-rel-gen}c)\cr
2154: }$$
2155: 	These should be compared with their nonrelativistic
2156: counterparts \eqsref{isotropic} and (\ref{iso-gen}).
2157: 
2158: In the relativistic limit, the Maxwellian distribution \eqref{max}
2159: becomes \cite{DeGroot} 
2160: 	$$f_{em}(p)={n_e\over4\pi m_e^2cT_eK_2(\Theta^{-1})}
2161: \exp\biggl(-{{\cal E}\over T_e}\biggr),\eqn(\eqlab{rel-maxwell})$$
2162: 	where
2163: 	$${\cal E}=m_ec^2\gamma_e$$
2164: is the total electron energy,
2165: 	$$\Theta=T_e/m_ec^2=T_e/511\,\rm keV,$$
2166: and $K_n$ is the $n$th-order modified Bessel function of the
2167: second kind.  If we substitute $f_e^{(0)}(p)= f_{em}(p)$ into
2168: \eqsref{iso-rel-gen}, we obtain $F_{cp}^{e/e}/D_{cpp}^{e/e} =-v_e/T_e$.
2169: Thus we find that $f_{em}$ annihilates the electron-electron collision
2170: term $C(f_{em},\linebreak[0]f_{em})=0$.  The integrals in \eqref{iso-rel-gen} cannot
2171: be performed analytically with $f_e^{(0)}(p)= f_{em}(p)$ and so in the
2172: numerical code these are performed numerically.
2173: 
2174: For the Maxwellian distribution \eqref{rel-maxwell}, we define a
2175: thermal momentum
2176: 	$$p_{te}=\sqrt{m_eT_e},$$
2177: a thermal velocity
2178: $$v_{te}^2={1\over 3n_e}\int v_e^2
2179: f_{em}(p)\,\dvp={T_e\over m_e}\biggl(
2180: 1-{5\over2}\Theta+
2181: {55\over8}\Theta^2+\ldots\biggr),$$
2182: and a thermal collision frequency
2183: $$\nu_{te}={m_e\Gamma^{e/e}\over p_{te}^3}.$$
2184: 
2185: For $p\gg p_{te}$, the indefinite limits in the integrals
2186: in Eq.~(\ref{iso-rel-gen}) can be replaced by $\infty$, giving
2187: \cite{Mosher}
2188: $$\eqalignno{D_{cpp}^{e/e}
2189: &=\Gamma^{e/e} {v_{te}^2\over v_e^3},&(\eqlab{rel-high}a)\cr
2190: D_{c\theta\theta}^{e/e}
2191: &=\Gamma^{e/e}{1\over 2v_e}\biggl(1-{v_{te}^2\over v_e^2}\biggr),
2192: &(\ref{rel-high}b)\cr
2193: F_{cp}^{e/e}&=-\Gamma^{e/e} {v_{te}^2\over T_ev_e^2}.&(\ref{rel-high}c)\cr
2194: }$$
2195: 	These should be compared with \eqsref{iso-high}.
2196: 
2197: For a background which consists of just the first Legendre harmonic,
2198: the collision term is $C\bigl(f_{em}(p),\penalty100 f_e^{(1)}(p)
2199: \cos\theta\bigr)$.  This is given by \cite{relpap}
2200: 	$$\eqalignno{{C\bigl(f_{em}(p), f_e^{(1)}(p)\cos\theta\bigr)
2201: \over f_{em}(p)\cos\theta}\kern-7em&\kern7em
2202: ={4\pi \Gamma^{e/e}\over n_e}\times \cr
2203: \Biggl\{&{m_e f_e^{(1)}(p) \over \gamma_e}
2204: +{1\over 5}\int_0^p p'{}^2f_e^{(1)}(p'){m_e\over T_e}\biggl[
2205: {\gamma_e\over p^2}{v_e'\over\gamma_e'{}^3}
2206: \biggl({T_e\over m_ec^2}(4\gamma_e'{}^2+6)
2207: -{1\over3}(4\gamma_e'{}^3-9\gamma_e')\!\biggl)\cr
2208: &\qquad\qquad\qquad\qquad\qquad\qquad\qquad
2209: {}+{\gamma_e^2\over p^2}{v_e'\over\gamma_e'{}^3}
2210: \biggl({m_ev_e'{}^2\over T_e}\gamma_e'{}^3
2211: -{1\over3}(4\gamma_e'{}^2+6)\!\biggl)
2212: \biggr]\,dp'\cr
2213: &\qquad\qquad{}+{1\over 5}\int_p^\infty p'{}^2f_e^{(1)}(p'){m_e\over T_e}\biggl[
2214: {\gamma_e'\over p'{}^2}{v_e\over\gamma_e^3}
2215: \biggl({T_e\over m_ec^2}(4\gamma_e^2+6)
2216: -{1\over3}(4\gamma_e^3-9\gamma_e)\!\biggl)\cr
2217: &\qquad\qquad\qquad\qquad\qquad\qquad\qquad
2218: {}+{\gamma_e'{}^2\over p'{}^2}{v_e\over\gamma_e^3}
2219: \biggl({m_ev_e^2\over T_e}\gamma_e^3
2220: -{1\over3}(4\gamma_e^2+6)\!\biggl)
2221: \biggr]\,dp'\Biggr\}.&(\eqlab{rel-first-lin})\cr}$$
2222: 	[Compare with \eqref{first-lin}.]  The general solution of the
2223: linearized electron-electron collision operator $C(f_e,f_{em}) +
2224: C(f_{em},f_e)=0$ is
2225: 	$$f_e=(a+{\bf b}\cdot{\bf p}+d{\cal E})f_{em},$$
2226: where $a$, $\bf b$, and $d$ are arbitrary constants.  With $a=d=0$ and
2227: ${\bf b}=\hat{\bf p}_\parallel$, this provides a
2228: useful check on Eqs.~(\ref{iso-rel-gen}) and (\ref{rel-first-lin}) and their
2229: computational realizations.
2230: 
2231: In the example we show below, we use the electron-ion collision
2232: operator given by \eqref{coll-rel-ions} and the relativistic
2233: generalization of the truncated collision operator \eqref{truncated}
2234: 	$$C_{\rm trunc}^{e/e}\bigl(f_e({\bf p})\bigr)=
2235: C\bigl(f_e({\bf p}),f_{em}(p)\bigr)
2236: +C\bigl(f_{em}(p), f_e^{(1)}(p)\cos\theta\bigr),
2237: \Eqlab{rel-truncated}$$
2238: 	where the first term is given by \eqsref{isotropic-rel} and
2239: (\ref{iso-rel-gen}) and the second term by \eqref{rel-first-lin}.
2240: 
2241: \subsection{Wave-particle interaction}
2242: 
2243: We saw in \secref{ql} that the quasilinear diffusion operator had two
2244: principal ingredients:  the wave-particle resonance condition, and the
2245: diffusion paths.  Both of these are modified by relativistic effects.
2246: 
2247: The wave-particle resonance condition becomes
2248: 	$$\omega-k_\parallel v_{e\parallel}-n\Omega_e/\gamma_e=0,$$
2249: where $\Omega_e=q_eB/m_e$ is the rest-mass cyclotron frequency.
2250: Translating this into momentum space gives
2251: 	$$\omega\sqrt{1+p^2/m_e^2c^2}-k_\parallel p_\parallel-n\Omega_e=0.$$
2252: 	This modification of the resonance condition is important in
2253: the consideration of current drive by electron cyclotron waves
2254: \cite{Cairns}.
2255: 
2256: The diffusion paths are again given by surfaces of constant energy in
2257: the wave frame.  The expression for the energy in a frame moving at
2258: $(\omega/k_\parallel){\bf \hat p}_\parallel$ is 
2259: 	$${\cal E}'={{\cal E}-(\omega/k_\parallel)p_\parallel
2260: \over\sqrt{1-\omega^2/k_\parallel^2c^2}}.$$
2261: The diffusion paths are, therefore, given by
2262: 	$${\cal E}-(\omega/k_\parallel)p_\parallel=\rm const.$$
2263: 	These paths are parallel to the vector
2264: 	$$\biggl({\omega\over k_\parallel}-v_{e\parallel}\biggr)
2265: {\bf \hat p}_\perp+v_{e\perp}{\bf \hat p}_\parallel.$$
2266: 	This should be compared with the vector ${\bf a}_n$ defined in
2267: \secref{ql}.  The paths are ellipses or hyperbolae in momentum space
2268: depending on whether $\omega/k_\parallel$ is less than or greater than
2269: $c$ \cite{Karney-ec}.
2270: 
2271: For lower hybrid waves, we have $n=0$ and the diffusion is
2272: in the parallel direction.  We, therefore, generalize \eqsref{d-ql}
2273: by incorporating the modified resonance condition to read
2274: 	$$\mat D_w=D_w(p_\perp,p_\parallel)
2275: {\bf\hat p}_\parallel{\bf\hat p}_\parallel,\eqn(\eqlab{d-ql-rel}a)$$
2276: 	where
2277: 	$$D_w(p_\perp,p_\parallel)=\cases{
2278: D_0,&for $v_1<p_\parallel/\gamma_e<v_2$,\cr
2279: 0,&otherwise.\cr}\eqn(\ref{d-ql-rel}b)$$
2280: 
2281: \subsection{Example}
2282: 
2283: To illustrate the relativistic effects we show in \figref{rel-fig} the
2284: steady-state distribution function obtained by integrating the
2285: Fokker--Planck equation with electron-electron collisions given by
2286: $C_{\rm trunc}^{e/e}$ \eqref{rel-truncated} and electron-ion collisions
2287: given by $C^{e/i}$ \eqref{coll-rel-ions} with $Z_i=1$.  The quasilinear
2288: diffusion term is given by \eqref{d-ql-rel} with $D_0=1$, $v_1=0.4c$,
2289: and $v_2=0.7c$.  (Except for the perpendicular profile of $\mat D_w$,
2290: this is the same as the example given in \refref{relpap}.)  The
2291: integration is carried out with $M=N=100$ and $p_{\rm max}=20$.  We
2292: normalize all momenta to $p_{te}$, velocities to $p_{te}/m_e$ ({\sl
2293: not} $v_{te}$), the current density to $n_eq_ep_{te}/m_e$, the power
2294: density to $n_ep_{te}^2\nu_{te}/m_e$, etc.  Again we are principally
2295: interested in the current and the power dissipated.  These are defined
2296: by
2297: 	$$\eqalign{J&=
2298: {\numint (v_e\cos\theta)\over n},\cr
2299: P&={1\over n}\sum_{i=0}^{M-1}\sum_{j=0}^{N}
2300: 2\pi\sin\theta_{i+1/2} v_jp_j^2 S_{wp,i+1/2,j} \,\Delta p\Delta\theta,\cr}$$
2301: 	where $\numint$ is the generalization of \eqref{num-int} to
2302: momentum space, $n=\numint 1$.  [Compare these expressions with
2303: \eqsref{curr-def-grid} and (\ref{pd-def-grid}).]  In the steady state,
2304: we find $J=3.732\times10^{-3}$, $P=1.256\times10^{-4}$, and
2305: $J/P=29.72$.
2306: 
2307: Again a useful benchmark is provided by the electrical conductivity.  In
2308: the limit $E\rightarrow0$ this is correctly given if $C_{\rm
2309: trunc}^{e/e}$ is employed.  With $E=10^{-3}$, $Z_i=1$, $\Theta=0.01$,
2310: $M=N=100$, and $p_{\rm max}=10$, we find $J/E=7.307$, which differs from
2311: the true value of 7.291 by about $0.2\%$.  Values of the conductivity
2312: for various values of $Z_i$ and $\Theta$ are tabulated in
2313: \tabref{rel-conduct}.
2314: 
2315: \section{Adjoint Method}\label{adjoint}
2316: 
2317: \subsection{Introduction and example}
2318: 
2319: We have considered here techniques for solving the Fokker--Planck
2320: equation with an added quasilinear diffusion term.  This tends to be an
2321: expensive operation because the addition of the quasilinear diffusion
2322: term greatly increases the parameter space to be scanned.  For example,
2323: the study of lower hybrid current drive \cite{Karney-lh} included the
2324: results of some 50 runs with different values of $v_1$ and $v_2$.  Even
2325: so, no systematic study was made of the dependence on the
2326: parameters $D_0$ and $Z_i$.
2327: 
2328: However, the amount of work can be drastically reduced using the
2329: adjoint method.  This was introduced by Hirshman \cite{Hirshman} for the
2330: study of beam-driven currents.  Later, Antonsen and Chu \cite{Antonsen}
2331: used it to study rf-driven currents.
2332: 
2333: To illustrate the method, we will outline the analysis given by
2334: Antonsen and Chu \cite{Antonsen}.  The method begins by assuming that
2335: $f_e$ is close to a Maxwellian $f_{em}$ so that the linearized
2336: electron-electron collision operator $C_{\rm lin}^{e/e}$ \eqref{linear}
2337: can be used.  The quasilinear diffusion term is taken as a {\sl given}.  As
2338: pointed out in \secref{ql}, the Fokker--Planck equation then becomes an
2339: inhomogeneous equation, whose linear operator is independent of the
2340: wave drive.  Two further assumptions are made, namely that ${\bf E}=0$
2341: and that a steady state has been reached.  (Neither of these assumptions
2342: is necessary and they have been relaxed in \refref{Fischa}.)  The
2343: Fokker--Planck equation is then
2344: 	$$\eqalignno{C\bigl(f_e({\bf v})\bigr)&
2345: \equiv C_{\rm lin}^{e/e}\bigl(f_e({\bf v})\bigr)
2346: +C^{e/i}\bigl(f_e({\bf v})\bigr)\cr
2347: &=\nabla\cdot {\bf S}_w+ \biggl({m_ev^2\over
2348: 2T_e}-{3\over 2}\biggr)f_{em}(v) {\partial \ln T_e\over \partial t},
2349: &(\eqlab{Chap-eq})\cr}$$
2350: 	where we have inserted the Chapman--Enskog--Braginskii energy
2351: loss term to ensure that \eqref{Chap-eq} has a solution (i.e., to
2352: ensure that the Fokker--Planck equation reaches a steady state).  Taking
2353: the energy moment of this equation, and noting that the collision
2354: operator is energy conserving, we find the equation for $\partial
2355: T_e/\partial t$
2356: 	$${\partial\over \partial t}\biggl({3\over2}n_eT_e\biggr)=P,$$
2357: 	where $P$ is given by \eqref{pd-def}.
2358: 
2359: The straightforward approach is now to solve \eqref{Chap-eq} for a
2360: particular ${\bf S}_w$, determine the electron distribution $f_e$, and hence
2361: find the rf-driven current.  The adjoint method gives a way of computing
2362: the current without having to find $f_e$.  Consider first the ``adjoint''
2363: problem 
2364: 	$$C\bigl(f_{em}(v)\chi({\bf v})\bigr)=-q_ev_\parallel f_{em}(v),
2365: \Eqlab{adjoint-eq}$$
2366: 	where we require that $f_{em}\chi$ have zero density and zero energy.
2367: This is the Spitzer--H\"arm equation for the perturbed distribution in
2368: the presence of an electric field ${\bf E}=T_e{\bf \hat v}_\parallel$.
2369: Let us multiply \eqref{adjoint-eq} by $f_e/f_{em}$ and integrate over
2370: velocity. This gives
2371: 	$$J=-\int (f_e/f_{em}) C(f_{em}\chi)\,\dv,$$
2372: where $J$ is the current carried by the electron distribution $f_e$.
2373: Now we utilize the self-adjointness of the linearized collision
2374: operator
2375: 	$$\int\psi C(f_{em}\chi)\,\dv=\int\chi C(f_{em}\psi)\,\dv,$$
2376: 	together with \eqref{Chap-eq} for $C(f_e)$ to give
2377: 	$$J=\int {\bf S}_w({\bf v})\cdot\nabla\chi({\bf v})\,\dv.
2378: \Eqlab{adjoint-curr}$$
2379: 	\Eqref{adjoint-curr} is the desired expression for the current.
2380: The quantity $\chi$ serves as the Green's function for the current $J$.
2381: The current drive efficiency is given by
2382: 	$${J\over P} ={\int {\bf S}_w\cdot\nabla\chi\,\dv\over \int
2383: m_e{\bf S}_w\cdot{\bf v}\,\dv}.\Eqlab{adjoint-eff}$$
2384: 
2385: \subsection{Solving the adjoint equation}
2386: 
2387: In order to apply this method, we must determine $\chi$ by solving
2388: \eqref{adjoint-eq}.  Because $\chi({\bf v})$ consists of only the first
2389: Legendre harmonic $\chi^{(1)}(v)\cos\theta $, this equation reduces to a
2390: one-dimensional integro-differential equation,
2391: 	$${1\over v^2}{\partial\over\partial v}
2392: v^2 D_{cvv}^{e/e}{\partial\chi^{(1)}\over\partial v}
2393: -{m_e v\over T_e}D_{cvv}^{e/e}{\partial\chi^{(1)}\over\partial v}
2394: -{2D_{c\theta\theta}^{e/e}+\Gamma^{e/e}Z_i/v\over v^2}\chi^{(1)}
2395: +I^{e/e}(\chi^{(1)})+q_ev=0,\eqn(\eqlab{adjoint1})$$
2396: 	where $D_{cvv}^{e/e}$ and $D_{c\theta\theta}^{e/e}$ are given by
2397: \eqsref{iso-max}, and $I^{e/e}$ is defined by
2398: 	$$I^{e/e}(\chi^{(1)}(v))=
2399: {C(f_{em}(v),f_{em}(v)\chi^{(1)}(v)\cos\theta )\over f_{em}(v)\cos\theta }$$
2400: 	[see \eqref{first-lin}].
2401: 
2402: In general, \eqref{adjoint1} must be solved numerically.  This is done
2403: by constructing the partial differential equation by setting the
2404: left-hand side of \eqref{adjoint1} equal to ${\partial
2405: \chi^{(1)}/\partial t}$.  The resulting equation is integrated in time
2406: with arbitrary initial conditions until a steady state is reached.  The
2407: integration is carried out in the domain $0<v<v_{\rm max}$ and the
2408: boundary conditions are taken to be $\chi^{(1)}(v=0)=0$ and
2409: $\partial^2\chi^{(1)}(v=v_{\rm max})/\partial v^2=0$.
2410: 
2411: Approach to the steady state is accelerated by treating the first three
2412: terms in \eqref{adjoint1} fully implicitly; i.e., in order to compute
2413: $[\chi^{(1)}(t+\Delta t)-\chi^{(1)}(t)]/\Delta t$, we evaluate these terms at
2414: $t+\Delta t$.  This means that very large time steps can be used.  The
2415: integral term $I^{e/e}(\chi^{(1)})$ is treated explicitly and is
2416: reevaluated at each time step.  The resulting difference equations form
2417: a tridiagonal matrix which can be solved by Gaussian elimination.
2418: 
2419: Because the adjoint equation is the same as the equation for the perturbed
2420: distribution in the presence of a weak electric field, we can solve
2421: \eqref{adjoint1} to obtain values of the electrical conductivity which
2422: is defined by
2423: 	$${J\over E}=\int {q_e v_\parallel\over T_e}f_{em}(v)\chi({\bf v})\,\dv
2424: ={4\pi q_e\over3T_e}\int v^3 f_{em}(v)\chi^{(1)}(v)\,dv.$$
2425: 	This procedure was carried out using the method outlined above
2426: with $v_{\rm max}=15 v_{te}$, $\Delta v=0.001 v_{te}$, and $\Delta
2427: t=1000/\nu_{te}$.  Because we are only working with a one-dimensional
2428: equation, it is possible to use a much finer mesh than with
2429: two-dimensional problems and so obtain results which are effectively
2430: ``exact.''  The results are summarized in \tabref{conduct} where we have
2431: also included the results from use of approximate collision operators.
2432: The same technique is easily generalized to relativistic plasmas using
2433: the collision operator given in \secref{relativity}.  This gives the
2434: relativistic corrections to the conductivity which are given in
2435: \tabref{rel-conduct}.
2436: 
2437: When the adjoint method is applied to more complicated situations
2438: (e.g., including a dc electric field), a two-dimensional equation must
2439: be solved.  We can then use many of the techniques for the solution of
2440: the Fokker--Planck equation, which have been presented in the preceding
2441: sections.
2442: 
2443: \subsection{Discussion}
2444: 
2445: Let us assess the work involved in utilizing the adjoint method.  Once
2446: the adjoint equation has been solved, the current and the efficiency
2447: are immediately given in terms of ${\bf S}_w$ by \eqsref{adjoint-curr}
2448: and (\ref{adjoint-eff}).  Instead of having to solve the Fokker--Planck
2449: equation afresh for every form of ${\bf S}_w$, a couple of velocity
2450: integrals over ${\bf S}_w$ suffice to give the important quantities.
2451: The parameter space that must be scanned in order to give a complete
2452: understanding of the physics is greatly reduced.  The adjoint method
2453: does not give the electron distribution $f_e$ nor the rf-induced flux
2454: ${\bf S}_w$.  On the other hand, a crude estimate of ${\bf S}_w$ gives
2455: an accurate estimate of the efficiency because \eqref{adjoint-eff}
2456: involves the ratio of two integrals over ${\bf S}_w$.  An effective way
2457: to use this method within a ray-tracing code would be to determine
2458: ${\bf S}_w$ from a solution of the one-dimensional Fokker--Planck
2459: equation \cite{Fisch1} and to use this to determine both $J$ and $P$
2460: from \eqsref{adjoint-curr} and (\ref{pd-def}).  The code thereby
2461: benefits from an accurate determination of the current drive efficiency
2462: while the high computational costs of integrating the two-dimensional
2463: Fokker--Planck equation are avoided.
2464: 
2465: Because the current drive efficiency is determined by a single function
2466: $\chi$, it is possible to ask questions not readily answerable from
2467: numerical solutions of the Fokker--Planck equation.  Examples are: What
2468: is the asymptotic form for the efficiency as the wave phase velocity
2469: becomes large?  What is the maximum possible efficiency for a particular
2470: class of waves?
2471: 
2472: Besides determining the current, the adjoint method can be adapted to
2473: give other moments of the electron distribution by changing the
2474: right-hand side of \eqref{adjoint}.  This can then give, for example,
2475: the perpendicular energy of the electrons, bremsstrahlung radiation,
2476: etc.  This method has been used to determine the current-drive
2477: efficiency in a relativistic plasma \cite{relpap}.  Recent developments
2478: of the method \cite{Fischa} allow the determination of arbitrary
2479: moments of $f_e$ (not just the current $J$), and the determination of
2480: the time development of such moments.  These have been applied to the
2481: study of rf current ramp-up \cite{ramp-pap}.
2482: 
2483: \section{Conclusions}
2484: 
2485: In the last fifteen years, Fokker--Planck codes have gone from esoteric
2486: programs developed by a few researchers which could only be run on a
2487: few machines to widely available tools used by a large number of
2488: physicists on many different computers.  This has been due to the large
2489: increase in computer power available to the average physicist and the
2490: pioneering efforts of Killeen {\sl et al.}\ \cite{Killeen1,Killeen2}.
2491: 
2492: In this paper, I have given a detailed description of a particular
2493: implementation of a code to solve the Fokker--Planck equation with
2494: emphasis on a particular application, namely current drive by
2495: lower hybrid waves.  There are many other implementations of this code
2496: that have been applied to a large variety of interesting problems.
2497: My goal has been to illustrate the main numerical problems by means of
2498: concrete examples.  The methods presented here cover the major numerical
2499: problems that are encountered in all Fokker--Planck codes.
2500: 
2501: There are two areas which still require attention.  Firstly, improved
2502: methods for obtaining the steady-state solution of the Fokker--Planck
2503: equation are needed.  Here the multigrid method offers the best promise for
2504: substantial savings over the other methods described in this paper.
2505: Secondly, the adjoint methods outlined in \secref{adjoint} should be
2506: extended and applied to a wider range of problems.  Ray-tracing codes
2507: still need to be modified to accept the results of these calculations.
2508: 
2509: \section*{Acknowledgments}
2510: 
2511: I would like to thank N.  J.  Fisch for a very fruitful collaboration
2512: extending over several years on various problems in rf current drive,
2513: which provided the impetus for the work described here.
2514: 
2515: This work was supported by the United States Department of Energy under
2516: Contract DE--AC02--76--CHO--3073.
2517: 
2518: \appendix\section{Numerical Techniques}\label{numerical} \input fortran.sty
2519: 
2520: In this appendix, various fragments of code are shown.  A two-di\-men\-sion\-al
2521: Fokker--Planck code is ideally suited to a vector processing machine
2522: like the Cray--1.  However, care must be taken to order the loops
2523: correctly, otherwise they will not vectorize.
2524: 
2525: The first example is the computation of the current
2526: $\numint(v\cos\theta)$ \eqref{num-int}.  This illustrates the
2527: rather peculiar way in which {\sc FORTRAN} code must be written in
2528: order to take advantage of the Cray--1's architecture \cite{CFT}.  We
2529: assume that the arrays and variables given in \tabref{fortran-vars}
2530: have been initialized as indicated.
2531: 	$$\fortcode{
2532: &	\dimension temp(0:iy-1)		\cr
2533: &	\do 1 i=0,iy-1			\cr
2534: &\1	temp(i)=0.0			\cr
2535: 1&	\continue			\cr
2536: &	\do 3 j=0,jx-1			\cr
2537: &\1	\do 2 i=0,iy-1			\cr
2538: &\2	temp(i)=temp(i)+x(j)\^3*f(i,j)	\cr
2539: 2&\1	\continue			\cr
2540: 3&	\continue			\cr
2541: &	\do 4 i=0,iy-1			\cr
2542: &\1	temp(i)=sn(i)*cn(i)*temp(i)	\cr
2543: 4&	\continue			\cr
2544: &	cur=0.0				\cr
2545: &	\do 5 i=0,iy-1			\cr
2546: &\1	cur=cur+temp(i)			\cr
2547: 5&	\continue			\cr
2548: &	cur=2.0*pi*dx*dy*cur		\cr
2549: }$$
2550: 	The important point is that the inner loop (with label {\it 2})
2551: vectorizes.  This would not happen if the order of the
2552: loops were reversed.  There is no particular advantage in taking the
2553: computation of $\fort x(j)\^3$ out of the inner loop since the {\sc
2554: CFT} compiler does this automatically.  The only loop that the compiler
2555: treats inefficiently is the last one.  In fact, we replace this by a
2556: call to the {\sc OMNILIB} routine ssum.
2557: 
2558: The second example is computing the integral part of the
2559: truncated collision operator
2560: $C\bigl(f_m(v),\penalty100
2561: f^{(1)}(v)\cos\theta\bigr)/\cos\theta$ \eqref{first-lin}.  Here
2562: again it is easy to arrange so that most of the code vectorizes
2563: \cite{McCoy}.  The computation of this term is then relatively
2564: inexpensive compared with the other computations.
2565: 	$$\fortcode{
2566: &	\dimension {\it s0}(-1:jx-1),{\it s3}(0:jx),{\it s5}(0:jx),
2567: 				{\it f1}(0:jx-1)		\cr
2568: &	\do 1 j=0,jx-1						\cr
2569: &\1	{\it f1}(j)=0.0						\cr
2570: 1&	\continue						\cr
2571: &	\do 3 i=0,iy-1						\cr
2572: &\1	\do 2 j=0,jx-1						\cr
2573: &\2	{\it f1}(j)={\it f1}(j)+1.5*dy*sn(i)*cn(i)*f(i,j)	\cr
2574: 2&\1	\continue						\cr
2575: 3&	\continue						\cr
2576: &	\do 4 j=0,jx-1						\cr
2577: &\1	{\it s0}(j-1)=dx*{\it f1}(j)				\cr
2578: &\1	{\it s3}(j+1)={\it s0}(j-1)*x(j)\^3			\cr
2579: &\1	{\it s5}(j+1)={\it s3}(j+1)*x(j)\^2			\cr
2580: 4&	\continue						\cr
2581: &	{\it s3}(0)=0.5*{\it s3}(1)				\cr
2582: &	{\it s5}(0)=0.5*{\it s5}(1)				\cr
2583: &	\do 5 j=1,jx-1						\cr
2584: &\1	{\it s3}(j)={\it s3}(j-1)+0.5*({\it s3}(j)+{\it s3}(j+1))\cr
2585: &\1	{\it s5}(j)={\it s5}(j-1)+0.5*({\it s5}(j)+{\it s5}(j+1))\cr
2586: 5&	\continue						\cr
2587: &	{\it s0}(jx-1)=0.5*{\it s0}(jx-2)			\cr
2588: &	\do 6 j=jx-2,0,-1					\cr
2589: &\1	{\it s0}(j)={\it s0}(j+1)+0.5*({\it s0}(j)+{\it s0}(j-1))\cr
2590: 6&	\continue						\cr
2591: &	\do 7 j=0,jx-1						\cr
2592: &\1	{\it c1}(j)=({\it s5}(j)/5.0-{\it s3}(j)/3.0)/x(j)\^2	\cr
2593: \cont&\1\hphantom{{\it c1}(j)=}
2594: 	+{\it s0}(j)*(x(j)\^2/5.0-1.0/3.0)*x(j)			\cr
2595: &\1	{\it c1}(j)=4*pi*{\it fm}(j)*({\it f1}(j)+{\it c1}(j))\cr
2596: 7&	\continue						\cr
2597: }$$
2598: 	All the loops vectorize with the exception of the indefinite
2599: integration loops (with labels {\it 5} and {\it 6}).  Most of the time
2600: is spent in the inner loop {\it 2} during the computation of $f^{(1)}$
2601: \eqref{legend-decomp}.
2602: 
2603: Finally, we consider vectorized Gaussian elimination.  This subroutine
2604: performs Gaussian elimination for the tridiagonal system of equations
2605: 	$$x_{i,j}+ \half\Delta t
2606: (a_{i,j}x_{i-1,j}+b_{i,j}x_{i,j}+c_{i,j}x_{i+1,j})=y_{i,j},$$
2607: 	to give $x_{i,j}$ for $0\le i< n$, $0\le j<m$.  The coefficients
2608: satisfy $a_{0,j}=c_{n-1,j}=0$.  A substantial fraction of the running time of
2609: the Fokker--Planck code is spent in this subroutine.  When implemented
2610: for a single system of equations $m=1$, this leads to ``vector
2611: dependencies'' which inhibit vectorization.  The solution is to solve
2612: the $m$ systems in parallel with $j$ being the index for the inner
2613: loops.  In the subroutine below, it is assumed that all the matrices are
2614: the same size, that the spacing in memory between $x_{i,j}$ and
2615: $x_{i+1,j}$ (the solution direction) is $ns$, and that the spacing
2616: between $x_{i,j}$ and $x_{i,j+1}$ (the vectorizing direction) is $ms$.
2617: This subroutine uses $x$ and $y$ as temporary storage; thus the initial
2618: data in $y$ are destroyed.
2619: 	$$\fortcode{
2620: &	\subroutine {solve}(x,ns,n,ms,m,a,b,c,y,dt)		\cr
2621: &	\dimension  x(0:ms-1,0:m-1),y(0:ms-1,0:m-1),		\cr
2622: \cont&	\hphantom{\dimension}a(0:ms-1,0:m-1),b(0:ms-1,0:m-1),	\cr
2623: \cont&	\hphantom{\dimension}c(0:ms-1,0:m-1)			\cr
2624: &	{\it dt2}=0.5*dt					\cr
2625: &	\do 2 i=0,n-1						\cr
2626: &\1	ia=ns*(i-1)						\cr
2627: &\1	ib=ns*i							\cr
2628: &\1	\do 1 j=0,m-1						\cr
2629: &\2	den=1.0/(1.0+{\it dt2}*(b(ib,j)+a(ib,j)*y(ia,j)))	\cr
2630: &\2	x(ib,j)=(y(ib,j)-{\it dt2}*a(ib,j)*x(ia,j))*den		\cr
2631: &\2	y(ib,j)=-{\it dt2}*c(ib,j)*den				\cr
2632: 1&\1	\continue						\cr
2633: 2&	\continue						\cr
2634: &	\do 4 i=n-2,0,-1					\cr
2635: &\1	ib=ns*i							\cr
2636: &\1	ic=ns*(i+1)						\cr
2637: &\1	\do 3 j=0,m-1						\cr
2638: &\2	x(ib,j)=y(ib,j)*x(ic,j)+x(ib,j)				\cr
2639: 3&\1	\continue						\cr
2640: 4&	\continue						\cr
2641: &	\return							\cr
2642: &	\end							\cr
2643: }$$
2644: 	There are a couple of tricky points here.  Firstly, we use
2645: nonstandard indexing into the arrays.  The element $x_{i,j}$ is accessed
2646: by the array element $x(ns*i,j)$.  If $ms=1$, then $ns*i$ will generally
2647: exceed the upper bound $ms-1$ on the first dimension of the arrays.
2648: This type of array indexing may cause problems with compilers that
2649: perform bounds checking.  Secondly, we have utilized the fact that
2650: $a(0)=c(n-1)=0$ and assumed that an arbitrary (possibly undefined)
2651: number multiplied by zero will give zero.  If this is not the case, the
2652: $i=0$ and $i=n-1$ iterations in the loop with label $\it 2$ will have
2653: to be split off from the rest of the loop and treated separately.
2654: 
2655: This subroutine is sufficiently general to be used for both the
2656: matrix inversions required in implementing \eqref{split}.  Assuming
2657: that all the matrices are dimensioned by, for example,
2658: 	$$\fortcode{&\dimension f(0:iyl-1,0:jxl-1)\cr}$$
2659: 	then the inversions are obtained by
2660: 	$$\fortcode{
2661: &	\call {solve}(xia,iyl,jx,1,iy,ax,bx,cx,phi,dt)\cr
2662: &	\call {solve}(xib,1,iy,iyl,jx,ay,by,cy,xia,dt)\cr}$$
2663: 
2664: \vspace{0.5in}
2665: \bibliographystyle{aip}\bibliography{fp}
2666: 
2667: \begin{thetables}{99}
2668: \newdimen\digitwidth\setbox0=\hbox{\rm0}\digitwidth=\wd0
2669: \newdimen\minuswidth\setbox0=\hbox{\mathsurround=0pt$-$}\minuswidth=\wd0
2670: {\catcode`?=\active\catcode`+=\active
2671:  \gdef\spdef{\offinterlineskip
2672:  \catcode`?=\active\def?{\kern\digitwidth}%
2673:  \catcode`+=\active\def+{\mathbin{\hbox to\minuswidth{\hfil}}}}}
2674: 	\tableitem{conduct} The electrical conductivity for various
2675: values of the ion charge $Z_i$ and for various electron-electron
2676: collision operators.  The conductivities are normalized to
2677: $n_eq_e^2/m_e\nu_{te}$.
2678: 	$$\vcenter{\tabskip.5em\spdef
2679: \halign{\strut\hfil\rm #\hfil\tabskip2em&\hfil$# $\hfil
2680: &\hfil$# $\hfil&\hfil$# $\hfil&\hfil$# $\hfil
2681: \tabskip.5em\cr\noalign{\hrule\vskip1.5pt\hrule}
2682: Collision operator&Z_i=1&Z_i=2&Z_i=5&Z_i=10\cr
2683: \noalign{\hrule}
2684: linearized        &   7.429  &   4.377  &   2.078  &   1.133  \cr
2685: drifting          &   6.331  &   3.876  &   1.932  &   1.084  \cr
2686: Maxwellian        &   3.773  &   2.824  &   1.660  &   0.998  \cr
2687: high-velocity     &   2.837  &   2.310  &   1.489  &   0.938  \cr
2688: \noalign{\hrule\vskip1.5pt\hrule}}}$$
2689: 	\tableitem{rel-conduct} The electrical conductivity of a
2690: relativistic plasma for various values of the ion charge $Z_i$ and for
2691: various electron temperatures.  The conductivities are normalized to
2692: $n_eq_e^2/m_e\nu_{te}$ and the electron temperatures are given in terms
2693: of $\Theta=T_e/m_ec^2$.
2694: 	$$\vcenter{\tabskip.5em\spdef
2695: \halign{\strut\hfil$# $\hfil\tabskip2em&\hfil$# $\hfil
2696: &\hfil$# $\hfil&\hfil$# $\hfil&\hfil$# $\hfil
2697: \tabskip.5em\cr\noalign{\hrule\vskip1.5pt\hrule}
2698: \Theta&Z_i=1&Z_i=2&Z_i=5&Z_i=10\cr
2699: \noalign{\hrule}
2700: 0.0?      &   7.429  &   4.377  &   2.078  &   1.133  \cr
2701: 0.01      &   7.291  &   4.275  &   2.019  &   1.097  \cr
2702: 0.02      &   7.160  &   4.180  &   1.963  &   1.064  \cr
2703: 0.05      &   6.807  &   3.928  &   1.821  &   0.979  \cr
2704: 0.1?      &   6.317  &   3.590  &   1.636  &   0.872  \cr
2705: 0.2?      &   5.575  &   3.102  &   1.383  &   0.729  \cr
2706: \noalign{\hrule\vskip1.5pt\hrule}}}$$
2707: 	\tableitem{fortran-vars}  Meaning of {\sc FORTRAN}
2708: variables and arrays.
2709: 	$$\vcenter{\tabskip.5em
2710: \halign{\strut$# $\hfil\tabskip2em&$# $\hfil
2711: \tabskip.5em\cr\noalign{\hrule\vskip1.5pt\hrule}
2712: \hbox{\rm{\sc FORTRAN} name}&\hbox{\rm meaning}\cr
2713: \noalign{\hrule}
2714: dx&\Delta v\cr
2715: dy&\Delta\theta\cr
2716: dt&\Delta t\cr
2717: jx&N\cr
2718: iy&M\cr
2719: xg(j)&v_j\cr
2720: x(i)&v_{j+1/2}\cr
2721: yg(i)&\theta_i\cr
2722: y(i)&\theta_{i+1/2}\cr
2723: cg(i)&\cos\theta_i\cr
2724: cn(i)&\cos\theta_{i+1/2}\cr
2725: sg(i)&\sin\theta_i\cr
2726: sn(i)&\sin\theta_{i+1/2}\cr
2727: pi&\pi\cr
2728: f(i,j)&f_{i+1/2,j+1/2}\cr
2729: {\it fm}(j)&f_{m,j+1/2}\cr
2730: cur&\numint (v\cos\theta)\cr
2731: {\it f1}(j)&f^{(1)}(v_{j+1/2})\cr
2732: {\it c1}(j)*cn(i)&\left.C\bigl(f_m(v),f^{(1)}(v)\cos\theta\bigr)\right|_{i+1/2,j+1/2}\cr
2733: ax(i,j)&a_{v,i+1/2,j+1/2}\cr
2734: ay(i,j)&a_{\theta,i+1/2,j+1/2}\cr
2735: phi(i,j)&\phi_{i+1/2,j+1/2}\cr
2736: \noalign{\hrule\vskip1.5pt\hrule}}}$$
2737: \end{thetables}
2738: 
2739: \begin{thefigures}{99}
2740: 
2741: \figitem{coord-fig}{4in} The cylindrical and spherical coordinate systems.
2742: 
2743: \figitem{diff-fig}{3.5in} The relation between the resonance condition for
2744: quasilinear diffusion $\omega-k_\parallel v_\parallel-n\Omega_e=0$ and
2745: the diffusion path $({\bf v}-(\omega/k_\parallel){\bf\hat
2746: v}_\parallel)^2={\rm const}$.
2747: 
2748: \figitem{grid-fig}{3.5in} The numerical grid showing where the distribution
2749: function and the fluxes are defined.
2750: 
2751: \figitem{current}{4.5in} The current as a function of time for $Z_i=1$,
2752: $f(t=0)=f_m$, and rf diffusion given by \eqsref{d-ql} with $D_0=1$,
2753: $v_1=3$, and $v_2=5$.  Here we have $M=N=100$, $\Delta t=0.2$, and
2754: $v_{\rm max}=10$.  Electron-electron collisions are computed using
2755: $C_{\rm Max}^{e/e}$.
2756: 
2757: \figitem{f-steady}{4.5in} The steady-state distribution for the case shown in
2758: \figref{current}.  The contour levels are $f=(2\pi)^{-3/2}\times
2759: \exp[-\half(j/5)^2]$ for $j={\rm integer}$.  This gives equally spaced
2760: contours for a Maxwellian distribution with spacing $\delta v=\fract1/5$.
2761: The resonant region is shown.
2762: 
2763: \figitem{flux-fig}{4.5in} The flux plot for the case shown in
2764: \figref{f-steady}.  The plot was obtained by plotting contours of the
2765: stream function $A$, \eqref{stream-grid}.  The contour levels are
2766: $2\times10^{-5}(j+\half)$ for $j={\rm integer}$.
2767: 
2768: \figitem{conv-d}{4in} The current $J$ (a) and the efficiency $J/P$ (b) as
2769: functions of $\Delta v$.  The parameters are the same as for
2770: \figref{current} except that $M$ and $N$ are allowed to vary with
2771: $M=N$.  The plots show the results from runs with $N$ varying between
2772: 100 and 350 in steps of 5 and between 350 and 500 in steps of 50.
2773: 
2774: \figitem{trunc-flux}{4.5in} The flux plot when $C_{\rm trunc}^{e/e}$ is used.
2775: The parameters are otherwise the same as for \figref{flux-fig}.
2776: 
2777: \figitem{residu-fig}{5in} $R$ as a function of time when Chebyshev
2778: acceleration is applied to the example shown in \figref{f-steady}.  Here
2779: $1/\beta=0.05$, $1/\alpha=1000$, $K=20$, and $f(t=0)=f_m$.  The
2780: convergence criterion $R<10^{-9}$ is met after 400 steps at $t=790$.
2781: 
2782: \figitem{run-fig}{4.5in} The steady-state distribution in the presence of a dc
2783: electric field.  Here we have $Z_i=1$, $E=0.06$, $M=N=100$, $v_{\rm
2784: max}=10$ and electron-electron collisions are computed using $C_{\rm
2785: Max}^{e/e}$.  The contour levels are the same as for \figref{f-steady}.
2786: 
2787: \figitem{conv-e}{4in} The runaway rate $\gamma$ (a) and the current $J$ (b)
2788: as functions of $\Delta v$.  The parameters are the same as for
2789: \figref{run-fig} except that $M$ and $N$ are allowed to vary with
2790: $M=N$.  The plots show the results from runs with $N$ varying between
2791: 50 and 300 in steps of 10 and between 300 and 500 in steps of 50.
2792: 
2793: \figitem{run-flux}{4.5in} The flux plot for the runaway problem.  This
2794: illustrates the same case as shown in \figref{run-fig} except that a
2795: source of particles is introduced at the origin to balance the runaway
2796: loss $\gamma=5.148\times10^{-5}$.  One set of contour levels is
2797: $0.1\gamma(j+\half)$ for $j={\rm integer}$ and $-10\le j <10$ (these
2798: give the stream lines that run away and the outermost stream lines that
2799: encircle the central eddy).  The other set of contour levels is
2800: $2\times10^{-4}(j+\half)$ for $j={\rm integer}$ and $j>0$ (these are
2801: the innermost stream lines about the eddy).
2802: 
2803: \figitem{rel-fig}{4.5in} The steady-state distribution for $Z_i=1$,
2804: $\Theta=0.01$ ($T_e=5.11\,\rm keV$), and rf diffusion given by
2805: \eqref{d-ql-rel} with $D_0=1$, $v_1=0.4c$, and $v_2=0.7c$.  Here we have
2806: $M=N=100$ and $p_{\rm max}=20$.  Electron-electron collisions are
2807: computed using $C_{\rm trunc}^{e/e}$.  The contour levels are chosen to
2808: be $f=\sqrt\Theta \exp[-\sqrt{1+\Theta(j/3)^2}/\Theta]/[4\pi
2809: K_2(\Theta^{-1})]$ for $j=\rm integer$ which give equally spaced
2810: contours for a relativistic Maxwellian with spacing $\delta
2811: p=\fract1/3$.  [For $\Theta=0.01$ we have
2812: $K_2(\Theta^{-1})=1.019\sqrt{\pi/2}\sqrt\Theta\exp(-1/\Theta)$.]  The
2813: resonant region is shown.
2814: 
2815: \end{thefigures}
2816: \end{document}
2817: