Plasma Physics - B. Samuel Tanenbaum
Plasma Physics - B. Samuel Tanenbaum
TRENT UNIVERSITY
LIBRARY
PLASMA PHYSICS
Digitized by the Internet Archive
in 2019 with funding from
Kahle/Austin Foundation
https://archive.org/details/plasmaphysicsOOOOtane
McGraw-Hill Physical and Quantum Electronics Series
PLASMA PHYSICS
Copyright © 1967 by McGraw-Hill, Inc. All
Rights Reserved. Printed in the United States
of America. No part of this publication may be
reproduced, stored in a retrieval system, or
transmitted, in any form or by any means,
electronic, mechanical, photocopying, record¬
ing, or otherwise, without the prior written per¬
mission of the publisher.
Library of Congress Catalog Card Number 67-22971
62812
12 34567890 MAM M 7432106987
Dedicated to my parents
Rena A. Tanenbaum and Harry M. Tanenbaum
and to my wife
Carol Binder Tanenbaum
PREFACE
This text on plasma physics is written for students at the senior or first-year
graduate level. It is an extended version of notes for a first-year graduate
course in plasma dynamics that has been offered for many years at Case. The
book (like the course for which it was developed) is intended to serve both as
an introduction to the theory of plasmas for students planning to do further
work in the subject and as a survey for students with other research interests
who want to have some knowledge of plasmas.
The book has been written with the realization that a first course in
plasma physics may attract students with a wide variety of backgrounds.
Students taking the course at Case have included physicists, chemists, mathe¬
maticians, astronomers, mechanical engineers, and electrical engineers. For
this reason every effort has been made to make the book self-contained by
including brief developments of all the mathematics, electromagnetic theory,
fluid mechanics, and kinetic theory used in the text. In addition the material
is arranged so that, for the most part, the more elementary topics (both
conceptually and mathematically) are presented in the first half of the text,
and the more advanced topics are left for the final chapters. Hence an
undergraduate or a graduate student with a weak background in mathematics
should be able to read the first three chapters with no trouble, as well as a good
number of the more advanced topics in the final three chapters and in the
Appendix.
The topics included in the book clearly fail to cover all areas of plasma
physics. Aside from the first two introductory sections there is almost no
discussion of the experimental aspects of the subject. Moreover, there are
some important theoretical topics, like radiation and atomic processes in a
plasma, which are scarcely mentioned, and others, like plasma instabilities and
magnetohydrodynamics, which are covered very briefly. I have attempted,
however, to cover a broad selection of topics with particular emphasis on the
various models that are used to predict the behavior of plasmas. In addition
I have tried to introduce the reader to much of the common terminology and to
most of the “well-known results” in plasma physics. In this connection I
have attempted throughout to use derivations which do not require too much
mathematical sophistication and which are complete enough for easy recogni¬
tion of all the assumptions and approximations used.
The idea of developing this material into a text was suggested to me by
Professor 0. K. Mawardi, who heads the plasma research program at Case.
During the past several years, Professor Mawardi has developed a full sequence
of plasma-physics courses, which now includes one senior course, three regular
graduate courses, and a number of special-topic courses. His suggestions
concerning topics to include in the text were very valuable, as was his constant
encouragement.
Many of my other colleagues at Case have helped me in preparing this
text, either by reading critically portions of the manuscript or by helping me
to learn those areas of plasma physics where my own background was weak.
Those who were of particular help include Professors R. E. Collin, E. Reshotko,
W. B. Johnson, R. B. Block, E. J. Morgan, R. T. Compton, E. D. Thompson,
and L. S. Taylor. A number of other people have also suggested changes
which improved the text. I would especially like to thank Drs. D. A. Meskan,
S. P. Heims, L. H. Holway, Jr., G. A. Otteni, and E. L. Walker and several
anonymous reviewers chosen by the publisher. In addition, I am indebted
for many helpful comments to all the students who used earlier versions of this
material as course notes.
R. S. Hirsh, who kindly agreed to read through the complete manuscript,
corrected many small errors and made numerous suggestions which, I believe,
make the text easier for the reader to follow. His assistance is greatly
appreciated.
The number of books dealing with plasma physics and related topics has
grown rapidly in recent years. A list of many of these texts is included in the
Bibliography at the end of this book. Of these references, the ones which were
most useful to me are the fine texts “Physics of Fully Ionized Gases,” by
Spitzer, “Elements of Plasma Physics,” by Gartenhaus, and “Foundations of
Plasma Dynamics,” by Holt and Haskell; the excellent chapter by Mintzer in
“Mathematics of Physics and Chemistry,” vol. 2, edited by Margenau and
Murphy; and the very readable nonmathematical surveys “Project Sherwood:
The U.S. Program in Controlled Fusion,” by A. S. Bishop, and “Elementary
Plasma Physics,” by L. A. Arzimovich.
A number of additional acknowledgments are due here. First, Professor
R. E. Bolz, head of the Engineering Division, both by his encouragement and
by his kindness in providing a relatively light teaching load, made it possible
for me to complete this manuscript in a fairly short time. During this time
my work was supported in part by research grants to the plasma group from
the National Science Foundation and from the National Aeronautics and
Space Administration. Next, I am greatly indebted to Mrs. J. L. Griffen, who
did a remarkably accurate job of typing all the material involved in preparing
this book. Finally, I must express my gratitude to Professor D. Mintzer,
who has been teacher, fellow researcher, and friend for many years.
Despite the surprisingly large amount of work required, I have found the
preparation of this text to be a remarkable educational experience. I hope
that the students who use the book will find it useful and will learn from the
reading a sizable fraction of what I learned from the writing.
B. Samuel Tanenbaum
CONTENTS
PREFACE
1.1 Introduction 2
1.2 Applications of Plasma Physics 14
1.3 Review of Vector Analysis and Electromagnetic Theory 30
1.4 Motion of a Charge in Uniform Fields 36
1.5 Motion in a Nonuniform B Field 43
1.6 Adiabatic Invariants; Applications; Time-varying Fields 52
2 MAGNETOIONIC THEORY 61
BIBLIOGRAPHY 351
INDEX 355
chapter one
MOTION OF
A CHARGED
PARTICLE IN
ELECTROMAGNETIC
FIELDS
chapter one
1.1 INTRODUCTION
Definition of a plasma
In this text we are concerned with the behavior of plasmas. The term
plasma, first applied by Langmuir in 1928 to the ionized gas in an
electric discharge, is now used to describe a large class of basically
neutral mixtures containing some charged particles, which might be
positive or negative ions, free electrons, or holes (vacancies in the other¬
wise filled electron band of a solid that behave like electrons with a
positive charge). Although the most actively studied plasmas are still
ionized gases, some electrolytic liquids and some solids can also be
classified as plasmas.
Three basic features characterize plasmas and distinguish them
from ordinary solids, liquids, or gases. The most notable feature is
that some, at least, of the particles are charged and therefore interact
with each other in accordance with the coulomb-force law. This
coulomb force (which is proportional to R~2, with R the distance
between the particles) falls off very slowly with distance compared with
most other interparticle forces, f Hence, in a plasma every charged
particle interacts simultaneously with many of its neighbors, giving
the plasma a cohesiveness which is often compared to that of a mold or
jelly-
Debye shielding
f The force between most neutral molecules in a gas, for example, is large when
the particle centers are within one diameter of each other but very nearly zero
when they are farther apart.
1.1 INTRODUCTION 3
that
/ T\1/2
\d = 69.0 ( — j meters (1.1.1)
where T = temperature, °K
ne — electron number density, particles/m3
From the plot of \D shown in Fig. 1.1 it can be seen that the Debye
length is less than 1 cm for laboratory plasmas of all types, although
it rapidly becomes very large for gases with very low electron number
densities (or extremely high temperatures).
The Debye shielding effect, while characteristic of all plasmas, does
not occur in every mixture containing charged particles. Fora mixture
which can be treated with classical equations there are two necessary
conditions for shielding to occur. The first and most obvious require¬
ment is that the physical dimensions of the system be large compared to
\d, since otherwise there is simply 7iot enough room for the shielding to
occur. In addition, there must be enough electrons within a distance \d
from the disturbance to produce the shielding; hence the average dis¬
tance between electrons must be small compared with the Debye length,
6 8 10 12 14 16 18 20 22 24 26 28
Fig. 1.1 The Debye length (in meters) plotted as a function of electron number
density and temperature.
4 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
y3/2
Nd = |nc7rXz>3 = 1.37 X 106 ^ (1.1.2)
7le
From Fig. 1.2 it is clear that solids do not satisfy the plasma shielding
criterion ND » 1 if we use the usual lattice temperature (about 300°Iv
or less) in Eq. (1.1.2). However, because of the high electron number
densities in a solid, quantum-mechanical effects become important and
require the temperature of the electrons to be as much as three or four
orders of magnitude higher than the lattice temperature, thus bringing
some solids into the range where Nd exceeds 1.
To see how this comes about we note that, from the uncertainty
o
bo
o
ApAx>~ (1.1.3)
m(v2) _ (p2)
(1.1.4)
3k 3 niK
hne113
Ap > (1.1.5)
47r
h2n 2/3
T > -:0 ‘ - = 7.4 X 10_16ne2/3 (1.1.6)
487r2mK
temperature which the lattice itself could not attain without melting
or vaporizing. This value for the electron temperature, as can be seen
in Fig. 1.2, raises the value of Nd to values comparable to, or greater
than, 1; hence at least some shielding effects can occur.
Plasmas are sometimes defined to include all media that have some
charged particles and satisfy the shielding criteria, Nd > 1 and \d <
smallest linear dimension of the plasma. However, this definition is
really a little too broad, as can be seen by a simple example. From
Fig. 1.2 note that a gas with a temperature of 1000°Iv and with 106
electrons/m3 has a very large value for ND■ If nn, the number density
of neutral molecules, is also relatively low (as in the interstellar gas),
this medium should be treated as a plasma. However, if this is a
relatively dense gas (where nn is 15 or 20 orders of magnitude larger
than ne), then the effect of the electrons is negligible, and the medium
can be treated as a neutral gas.
The crucial difference between these two cases is that re„, the
average time an electron travels between collisions with neutral mole¬
cules, is large (perhaps days) for the rarefied gas and exceedingly small
(perhaps 10~u sec) for the dense one. Therefore, the electrons can
have a relatively independent behavior in the former case, having a
separate drift velocity and temperature, for example, from that of the
neutrals, while in the latter case there are so many electron-neutral
collisions per second that the electrons are always forced to be in com¬
plete equilibrium with the neutrals.
This shows up clearly in a calculation of the dc. electrical con¬
ductivity ao of a weakly ionized gas, where (in Sec. 2.2) we find that
do = nfi^Ten/m = 2.8 X 10~8ncTen. Thus, in the example we were just
considering, the interstellar gas is a very good conductor, while the
dense gas is a very poor one.
From the analysis of collision frequencies in Chap. 5 we find that,
for a gas
4 X 1015
(1.1.7)
nnTe1/2
so that for ionized gas in the laboratory ren is typically between 10-5
and 10-9 sec. In a solid, where electrons suffer collisions with vibrating
lattice ions, with impurities, and with lattice imperfections, t is much
shorter than this unless one works with very pure crystals and at a
low lattice temperature (where lattice vibrations are greatly reduced).
However, at temperatures of a few degrees Kelvin, using very pure
samples of bismuth, values of r in excess of 10-9 sec are possible.
1.1 INTRODUCTION 7
Although these values for r may seem short, one must remember that
plasma experiments often involve waves at microwave (or higher) fre¬
quencies, where the period of an oscillation is less than 10-10 sec; hence
the time between collisions is still large compared to the time during
which the plasma parameters are changing. For this reason, when
dealing with wave phenomena, if to is the frequency in radians per
second, then the product cor is large when plasma effects are important
and small when there are too many collisions for the electrons to be
uncoupled from the rest of the system.
Naturally occurring plasmas
Throughout- the bulk of this text we shall be concerned with two closely
related goals: (1) the development of the various sets of equations that
are used to study the behavior of plasmas, together with some discussion
of the limitations and approximations involved in each approach; and
(2) a few solutions to the plasma equations, chosen to illustrate either
important physical properties of the plasma or useful mathematical
methods for dealing with the equations.
In view of these goals we shall not devote much space to the many
actual and potential applications of plasma physics. However, it is
important to describe some of these briefly at least, since it is these
applications which motivate (and finance) most of the research in the
subject.
To start, our interest in plasmas results from the fact that most-
of the universe, with a few exceptions like the earth, is a plasma. First,
there are the stars, which because of their high temperatures are almost
completely ionized. Next there is the interstellar gas, which, despite
its low number density, still exhibits some plasma properties. Finally,
there are the numerous special plasmas that show up in a wide variety
of geophysical and ast-rophysical problems. These include the iono¬
sphere, the region of partly ionized gas that extends from about 50 km
above the earth to several earth radii;f the Van Allen belts of energetic
ionized particles trapped at high altitudes by the earth’s magnetic
field; the beams of plasma which erupt from the sun during magnetic
storms and which are related to the spectacular aurora borealis seen in
the earth’s upper atmosphere; the steady stream of plasma known as
the solar wind, which greatly distorts the earth’s magnetic field at
heights greater than two or three earth radii; and the spectacular gas
discharges created by a lightning stroke.
Solid and liquid plasmas
A great many solid, liquid, and gaseous plasmas, man-made as well as
natural, are present on earth. Among the solids alone, for example,
f One earth radius is about 6370 km.
8 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
there are several kinds of plasma. These can be divided into two
classes, compensated plasmas, consisting of electrons and holes, and
uncompensated plasmas, where the numbers of mobile negative and
positive charges are not equal and overall charge neutrality is main¬
tained by heavy positive or negative ions. Compensated plasmas
include metals like iron and tungsten, semimetals (so-called because
the number density of electrons and holes is much lower than for the
metals) like bismuth and antimony, and semiconductors. Uncom¬
pensated plasmas include metals like copper and sodium and semi¬
metals or semiconductors that are doped by adding impurity atoms
which release electrons or holes, depending on whether the impurity
atom has more or fewer electrons than the regular atoms of the lattice.
In addition to these “natural” solid-state plasmas, i.e., plasmas
which exist in the crystals as they are grown, it is also possible to
create plasmas in a semiconductor by applying an external electric field
strong enough to cause avalanche breakdown. This process occurs when
the electrons gain enough energy from the applied field to dislodge other
electrons when they collide with lattice atoms, so that a high electron
number density is quickly created.
Examples of liquid plasmas include the liquid metals, like mercury,
which was used in the early experimental work in magnetohydro¬
dynamics, and electrolytic solutions of all types. (The Debye shielding
theory, in fact, was first developed for electrolytes.)
Argon 15.68
Cesium 3.87
Hydrogen 13.53
Helium 24.46
Mercury 10.39
Potassium 4.32
Sodium 5.12
Oxygen 13.55
1.1 INTRODUCTION 9
a(%)
100
- 75
50
25
-3-2-1 0 1 2 3 T/U
r = Ce~UIT C» 1 (1.1.10)
U
T1/2 (1.1.11)
In C
Thus, as shown in Fig. 1.4, 50 percent of the atoms are in the upper
energy level when the temperature (in electron volts) is still well belotv
the excitation energy U.
Another interesting feature in this second case is that the percent
of particles in the excited state changes from nearly 0 to nearly 100
over a narrow temperature range, much as in an ordinary phase transition
from a solid to a liquid or from a liquid to a gas. A good measure of this
is the temperature difference AT that would exist between a = 0 and
a = 100 percent if the curve of a(T) were a straight line with the slope
of the true a(T) curve at T1/2. (Note the dotted curve in Fig. 1.4.)
From Fig. 1.4 it is easily seen that
Hence, from Eqs. (1.1.9) to (1.1.12) one easily finds that if C is assumed
Ar-w (LU3)
This result shows that if In C is sufficiently large, that is, ga » gb, the
curve of a(T) looks like a step function, with the transition from state
6 to state a occurring at 7\/2.
For the ionization problem that we are concerned with here, state a
is the ion-electron pair, state b is the neutral atom, and U is the ioniza¬
tion potential. Values for ga and gb are computed from quantum-
mechanical arguments^ and, in general, are quite complicated. How¬
ever, if the internal degrees of freedom of the particles are neglected,
one finds, to a good approximation, that
2-ir menT\31
C = g- = ni-
gb \ h2
2.405 X 1021T3/2nr1 T in °K
with rii the number of ions per cubic meter. When this expression is
used in Eq. (1.1.10), the result, known as the Saha equation, is clearly
Tm / dC\
6 [In (In C)]irm
C In C\df)Tut
J See, for example, W. B. Thompson, “An Introduction to Plasma Physics,”
pp. 17-20, Addison-Wesley Publishing Company, Inc., Reading, Mass., 1962.
12 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
Temperature, ev
Nonequilibrium plasmas
Thus far we have discussed only those gaseous plasmas produced by
raising the temperature to a point where the equilibrium percent ioniza¬
tion is reasonably high. Although this accounts for many astrophysical
f The effect is a little more complicated than Eq. (1.1.11) suggests, since In C
is a slowly varying function of temperature.
1.1 INTRODUCTION 13
ionosphere during the daylight hours, and when this radiation source is
“turned off” at sunset, the electron number density falls off gradually
to a much lower equilibrium value.
Lower-frequency optical or radio-frequency (r-f) radiation cannot
produce a plasma by photoionization, since each photon’s energy is
too small. However, a focused laser beam or a high-power r-f signal
has such a high electric-field strength that it can ionize a gas and produce
a plasma in the same manner as in an ordinary gas discharge. Ioniza¬
tion is also produced by x-rays or gamma rays, which have still higher
frequencies than ultraviolet. Although these high-energy sources are
not usually used to produce laboratory plasmas, their effect is utilized
in the ionization chamber, a device which determines the intensity of an
x-ray beam by measuring the number of ions created as the beam passes
through a gas in the chamber. In summary, there are a great many
ways to create a gaseous plasma. We have touched briefly on some of
the methods, and space prevents us from discussing still others, e.g.,
ionization produced by surface effects. For a full discussion of the
processes which produce gaseous plasmas the reader should consult one
of the many books dealing solely with this subject.f
Problem
1.1 Prepare a set of curves like Fig. 1.5 showing the percent ionization
of cesium as a function of temperature.
Having discussed the properties that define a plasma and some of the
many naturally occurring and man-made plasmas, we now want to note
some of the potential and (in some cases) existing practical uses for
plasmas. A quick list, chosen to give some idea of the wide range of
such applications, includes nuclear-fusion devices, magnetohydrody¬
namic (mhd) generators, plasma propulsion systems, thermionic con¬
verters, plasma amplifiers, gaseous lasers, arc jets, and fluorescent tubes.
Nuclear fusion
Among the items on this list the nuclear-fusion devices could have the
greatest impact from a practical standpoint, but they have also proved
to be the most difficult to develop. Nuclear fusion is the process
whereby two light nuclei combine to form a heavier one. In these
fusion reactions the number of protons and neutrons is conserved; but
when the initial nuclei are light ones (atomic mass number less than
20), the total final mass is a little less than the initial mass. Hence,
from Einstein’s familiar law, £ = me2, there is a large energy release.
Important examples of this are the reactions
The basic difficulty in achieving fusion is that the process requires the
interacting particles to approach within a distance of order ICE14 m.
(The nuclear “radius” is generally taken to be 1.5 X 1CE15.A1/3 m, with
A the atomic mass number.) However, the incident particles, being
positively charged, repel each other in accordance with the coulomb-
force law. In order to estimate the kinetic energy the particles must
have before they can overcome this coulomb barrier, note that the
potential energy due to the coulomb interaction between two nuclei
(with atomic numbers! Z\ and Z2) is
ZxZ2e2 (1.2.2)
47re0A
When the distance between the particles R is 1(E14 m, one finds that
Uc = 1-44 X 105ZiZ2ev. Since the average kinetic energy of the parti¬
cles is §T (with T in electron volts), it is clear that fusion interactions
f See A. S. Bishop, “Project Sherwood,” Addison-Wesley Publishing Com¬
pany, Inc., Reading, Mass., 1958.
J Note that, for a given nucleus, Z is the number of protons while A is the
total number of protons and neutrons.
16 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
will generally not occur in great numbers until the temperatures come
close to 105 ev or 109°K. (Actually, since at any temperature, some
of the particles have energies well above average, some fusion occurs at
temperatures as low as about 4 X 107°Iv.)
This is borne out in Fig. 1.6, where the cross section for each of
the fusion reactions shown in Eq. (1.2.1) is plotted as a function of
energy. As might be expected from Eq. (1.2.2), other fusion reactions
involving nuclei with larger Z values require still higher energies to
overcome the coulomb barrier (despite their slightly larger radii).
Thus, as a result of the coulomb barrier, fusion reactions occur only
when the interacting particles have energies that are at least in the
10-kev range, which means that the plasma must have temperatures on
the order of 100 million degrees Kelvin. (For this reason this is usually
called a thermonuclear process.)
10
£Z
o
3. Some care must be taken lest the energy loss due to radiation by
the fast-moving charged particles exceed the energy gain by fusion.
1. Cyclotron and other r-f heating schemes, in which the energy comes
from an electromagnetic wave, usually at some particular resonant
frequency of the plasma where (as we shall see) energy absorption
is very high.
2. Shock heating, in which the plasma is compressed and heated by
a shock wave.
3. Magnetic heating (also called adiabatic compression), in which the
energy of the particles is increased by slowly increasing a magnetic
field in the plasma. A special case is the pinch effect, where the
magnetic field is the self-induced field due to a current in the plasma.
Fig. 1./ Typical examples of the three basic confinement schemes, (a) Linear
pinch; (6) magnetic-mirror system ; (c) stellarator.
1.2 APPLICATIONS OF PLASMA PHYSICS 19
The nr criterion
For a practical device the confinement time r must be long enough
so that a significant percent of the fuel can undergo fusion. To estimate
how long this takes, consider the cylinder with cross-sectional area a,
where cr is the fusion cross section, and length vt, where v is the average
speed, v = (v2)112, of a nucleus. We may think of this cylinder as the
volume swept out by an average particle during the confinement time,
such that any other particle in this volume will undergo a fusion reaction
with the test particle. If n is the number density of the potential
collision partners, then the average number of particles in this volume
is 7iavr. Hence, to have a high probability of fusion we want
- ^ i (1.2.3)
ncrVT > 1
/3UrV/2 (1.2.4)
\ m J
f See L. A. Arzimovich, “Elementary Plasma Physics,” pp. 168-174, Blaisdell
Publishing Company, New York, 1965.
20 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
with m the nuclear mass. Hence if we assume that T — 300 kev, which
is about as high a temperature as seems feasible, then for the H2-H-
reaction, from Fig. 1.6 and Eq. (1.2.4), <r = 10~29 m2, and v = 6.5 X 106
m/sec so that the condition for a high probability of fusion [Eq.
(1.2.3)] becomes
Radiation losses
(1.2.8)
P0 = RfEf (1.2.9)
where Rf is the average number of fusion reactions per cubic meter per
second and Ef is the average energy release per reaction. To estimate
Rf note that, as in the discussion leading to Eq. (1.2.3), the probability
per unit time for a particle to have a fusion reaction is n,crv; hence, since
the number of particles per cubic meter is rq and each fusion involves
the reaction of two particles, if we neglect the statistical spread in the
velocities of the particles, we have to a good approximation
Rf = \n?av (1.2.10)
radiation and that, in addition, the radiation losses are a serious factor
in maintaining a self-sustaining fusion device. This is shown graphi¬
cally in Fig. 1.9, where PT is compared with Pg as a function of tempera¬
ture for both the H2-II2 and H2-II3 reactions. The crossover points
(where Pg exceeds Pr) are seen to be at about 4.5 kev for the H2-H3
reaction and 40 kev for the H2-H2 reaction.
Comparing the expression for Pg [Eqs. (1.2.9) and (1.2.10)] with
the bremsstrahlung-radiation loss shown in Eq. (1.2.7), we note that P0
is proportional to n,2 while Pr is proportional to Z2riine. Hence, for
a gas with no impurities, where ne = Znh the ratio of Pr to Pg is inde¬
pendent of ion number density, but if all else is equal, the radiation
loss is much more serious for the high-Z plasma than for the low. In
addition it is evident that, even for a hydrogen plasma, any high-Z
impurities will increase the bremsstrahlung radiation greatly. Thus,
most realistic fusion devices require the use of a hydrogen plasma of
such a high purity that this gas purification alone poses a major techno¬
logical problem, almost comparable to the heating and confinement
problems.
Problem
Mhd generators
— = —E = vB (1-2-13)
dz
Hence, upon integrating both sides from 0 to d, we find that the open-
circuit potential difference in volts between electrodes at z = 0 and
z = d is
<f>
I = (1.2.15)
Rl + Rp
&Rl
Pl (1.2.16)
(Rl + Rp)2
Fig. 1.10 Schematic diagrams for an mhd generator. Plasma flows in x direc¬
tion across magnetic field in y direction producing voltage <I> between electrodes
at z = 0 and z — d.
1.2 APPLICATIONS OF PLASMA PHYSICS 25
_d_
RP = (1.2.17)
Aa
Hence the maximum power that can be delivered to the load is easily
found, from Eqs. (1.2.14), (1.2.16), and (1.2.17), to be
d2
Rl = Rr = 1(E3 ohm (1.2.20)
p V (j
<t>
I = 5 X 104 amp (1.2.21)
2/0
Thus, in any realistic application there would be the need for some
conversion equipment to put this power into a more usable form.
then the actual velocity of the exhaust gas is v — u, with u{t) the
velocity of the rocket. Hence the force acting on the rocket is
— M(v — u), and from Newton’s law we have
or
u = (1.2.23)
» In (1.2.24)
with Uo and M0 the initial values of the velocity and mass of the rocket
and uf and Mf the final values. This result shows that a plasma
propulsion system (with v perhaps 100 times as large as in a chemical
rocket) can attain a given final velocity using a much smaller percent
of the mass for fuel than would be possible with a chemical rocket.
A second factor in the design of a rocket-propulsion system is the
ratio a of the final mass to the power output, that is,
Mf 2M i
a or M = (1.2.25)
iiu2 av‘
To introduce this factor into the expression for the final velocity, note
that if a rocket burns for a time t0, then
M o = Mf (1.2.26)
uf = v In ^1 + (1.2.27)
with V2 = 2to/a.
1.2 APPLICATIONS OF PLASMA PHYSICS 29
For a fixed value of V, uf has the form shown in Fig. 1.13, where
it can be seen that uf has a maximum of about 0.8F, which is obtained
when v = 0.5F. Note that under these optimal conditions, from
Eq. (1.2.26), Mo/Mf = 5. Hence 20 percent of the initial mass can
be used for the electric propulsion equipment and the payload.
An important practical factor involved in plasma rocket-propul¬
sion systems is that they achieve their high final velocity by means
of a small acceleration over a long time period. Hence, the force they
provide is much too modest to overcome the gravitational force close
to earth, and chemical rockets must still be used as the first stage of
any plasma propulsion system (see Prob. 1.3).
Problem
v(Mo - Mf)
-u(O)
M oto
In this section we shall introduce the basic mathematical tools and the
fundamental equations of electromagnetic theory needed ■ in our
analysis of plasmas.
Coordinate systems
f See, for example, Cruft Electronics Staff, “Electronic Circuits and Tubes,”
McGraw-Hill Book Company, New York, 1947.
1.3 REVIEW OF VECTOR ANALYSIS AND ELECTROMAGNETIC THEORY 31
Rectangular (x,y,z) dx dy dz
Cylindrical (r,<j>,z) dr r dct> dz
Spherical (R,8,4>) dR R d9 R sin 9 dcf>
3
A,a, = J A&i (1.3.3)
i= i
There are two types of products of vectors. The scalar, or dot, product
A • B = AiBi (1.3.4)
~ <i>
Fig. 1.14 Angles and lengths used in
cylindrical and spherical coordinates.
32 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
ai a2 a3
A X B = Ai A2 Az
Bz b2 Bz
Hence, if A and B are parallel, their cross product is zero, while if A and
B are perpendicular, their dot product is zero.
h iai h2 a2 hz&z
1 d d d
(1.3.9)
hih2h3 dui dlh duz
h\A i hiA 2 hzA
and
Problems
VXE=—B V x H = D + J
(1.3.12)
V • D = Pc V • B = 0
Pc = Xna9a (1.3.13)
a
and
J = lnaqaua d-3.14)
a
0 = pc + V • J (1.3.15)
The total charge contained in the volume is JpcdV. Hence, for the
volume element
J (pc + V • J) dV = 0 (1.3.18)
which can hold for all volumes only if Eq. (1.3.15) is true.
B = V x A (1.3.19)
1.3 REVIEW OF VECTOR ANALYSIS AND ELECTROMAGNETIC THEORY 35
V x E = —B = — V X A (1.3.20)
so that
V x (E + A) = 0 (1.3.21)
E + A - -V4> (1.3.22)
For many media I) and B are related to E and H by the simple con¬
stitutive equations
D = eE B = mH (1.3.23)
D = e • E (1-3.25)
Poynting’s theorem
To conclude this brief review of Maxwell’s equations, we shall derive
the Poynting theorem, which is an energy-conservation law for electro¬
magnetic fields. From the two Maxwell curl equations note first
that for a simple isotropic medium with n and e independent of time
This result is known as Poynting’s theorem, and the three terms all
have important physical interpretations. The one on the left is the
power flow into the volume (since —dS points into the volume); the
first term on the right is the rate of increase of the electric- and mag¬
netic-field energy in the volume (\tE2 and ^nIP are, respectively, the
electric- and magnetic-field energy densities); and the last term on
the right is the dissipative power loss in the volume. The vector on
the left, E x II, which is involved with the power flow of the electro¬
magnetic field, is called the Poynting vector.
The forces acting on such a particle are intimately related to the very
definitions of E and B. E is defined in terms of the force Fe produced
on a charge distribution placed in the field
Fe = JpcE dV (1.4.1)
Fe = 'EjpcdV = qE (1.4.2)
Fm = JJ x B dV = Jpcii X B dV (1.4.3)
Again, if the test charge is small, so that u and B are constant over the
volume where pc ^ 0, then
Fm = u X B\pcdV = qu X B (1-4.4)
From Eqs. (1.4.2) and (1.4.4) we conclude that the force on a point
charge (with velocity v) placed in an electromagnetic field is given by
F = q(E + v x B) (1-4.5)
From Newton’s law, if the Lorentz force is the only force acting on a
charged particle, we have
m ^ = g(E + v X B) (1.4.6)
Taking the dot product of v and each side of this equation, we have
d (1-4.7)
^mv2 qx • E
dt
38 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
or
This shows that the sum of the kinetic and potential energies stays
constant (with 4> the potential energy per unit charge).
We now consider the actual solution to Eq. (1.4.6) when E and B are
constants. Although this problem is rather simple, we shall treat it
in great detail, since many of the more complicated situations are
treated as perturbations from this problem. Choose the z axis to lie
in the B direction, so that B = Baz and E = Exax + Eyay + Eza,.
(In practice one can arrange this situation in the laboratory by wrap¬
ping a coil of wire around a tube containing a pair of large planar elec¬
trodes, as shown in Fig. 1.16.) The Lorentz equation (1.4.6) for this
case reduces to
Vz =
Vz — — E zt + vz(0) (1.4.11)
m
(1.4.12)
(1.4.13)
To find vy we substitute this result for vx directly into the first part of
Eq. (1.4.10), to have
Vy
1.1.11
E x B
Vd = (1.4.15)
B2
40 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
Note that this drift velocity is independent of q and m. Thus all the
charged particles will drift along together in this case, and the net
current flow due to the drift motion of a collection of charged particles
is [from Eq. (1.3.14y J
Thus, there is no drift current for the (usual) case when the net charge
density in the plasma is zero, i.e., when the total charge densities of
the positive and negatively charged particles balance out.
The actual solution for the path of the particle is obtained by integrat¬
ing each of these equations and applying the initial conditions on the
velocity and position of the particle. Let us choose our coordinates
so that initially we have
Y
</> = tan-1 — cod (1.4.20)
X
Z = V it
Thus the particle moves in a helical path. The radius |a| is called the
cyclotron radius or radius of gyration, and, as noted above,
F0 _ V0m
a (1.4.21)
w, qB
1.4 MOTION OF A CHARGE IN UNIFORM FIELDS 41
so that a increases with the mass and velocity of the particle but
decreases with the charge and the held strength. In addition, from
the expression for 0, note that $ increases with time for a negative
charge (wc < 0) and decreases with time for a positive charge. Hence,
as shown in Fig. 1.17, for an observer looking into the lines of a B held,
the positive charge spirals clockwise and the negative charge counter¬
clockwise. (Alternatively, if one’s thumb points in the direction of B,
the positive charge spirals like the fingers of the left hand and the nega¬
tive charge like the fingers of the right hand.)
When a perpendicular E held is present, the path of the particle is
obtained by integrating Eqs. (1.4.13) and (1.4.14). The exact orbit
depends strongly on the initial velocity. However, it is clear that the
path consists of both the spiraling cyclotron motion and the steady
drift in the direction perpendicular to E and B, as shown, for a typical
example in Fig. 1.17.
Problems
1.7 Plot the orbit for a particle when E = Eay and initially the
particle is located at the origin with zero velocity. Note the
difference in path for a positive and negative charge.
1.8 Plot the radius of gyration for an electron as a function of held
strength when F0 is the “thermal velocity” («T/m)1/2 and T has
the values 10-2, 1, 102, and 104 ev.
Fig. 1.17 Typical orbits for charged particles in crossed electric and
magnetic fields. Curves on the left are for E = 0.
42 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
= q{\ X B) + mg (1.4.22)
This is exactly like Eq. (1.4.6) for the motion in magnetic and electric
fields except for the change from the electric force qYL to the gravita¬
tional force mg. Thus, the gravitational force is entirely equivalent
to an electric field mg/q acting on the charge. In particular, where
the drift velocity for E and B fields was shown to be E X B/B2 we now
have for uniform g and B fields
_ m(g x B)
(1.4.23)
qB2
Note that for this case the drift is in opposite directions for electrons
and positive ions and is directly proportional to the mass. Thus the
current density produced by the g and B fields is
njiUjig X B) p(g X B)
Jo = ^ "Ab = ^ B2 B'2
(1.4.24)
SUMMARY
Problems
V • B = v X B = 0 (1.5.1)
B = V<hm (1.5.2)
with 4>m the magnetic scalar potential; while from the divergence condi¬
tion we have
V2<t>m = v • B = 0 (1.5.3)
Thus the magnetic field in which the particle moves is not completely
arbitrary but must, in fact, be the gradient of a solution to Laplace’s
equation.
As an example, suppose we let
where, to assure that B is still nearly uniform, we require that \az\ also
be small compared to unity. The x component of B we shall call the
curvature term', the ^-dependent part of Bz we call the gradient term.
Problem
1.11 Show that the following magnetic fields violate Maxwell’s equa¬
tions: B = B{z) az; B = B(x) a2.
cbk dT d4>
^ = ~aB0z (1.5.7)
dx aT = B"(1 + al) dz dx
Hence, the equation for each field line has the form
For the field line passing through the origin (x = z = 0), the
constant in Eq. (1.5.11) must be zero, so that near the origin
z2 = — (1 + iax) (1.5.12)
a a
where in the last step we use the assumption that |ax| « 1. We can
show that, to a good approximation, near the origin this is the arc of a
circle with radius a-1 and centered at x = a-1, z = 0. The equation
for such a circle is
(x — a-1)2 + z2 = or1
or, restricting attention to that part of the arc near the origin,
z2 = —— (1 tax)**— (1-5.13)
a a
Hence, as long as \ax\ <5C 1, this circle has the same equation as the
field line passing through the origin.
Problem
1.12 Verify that, to a good approximation, near the origin the equipo-
tential lines are radii of circles centered at x = a-1, z = 0 as shown
in Fig. 1.18.
vx — wcvy(l + olt)
qB o
VZ = —GJcOLZVy
m
X(0) = -a 7(0) = 0 Z( 0) = 0 a
a -l
Equipotentials
Field lines
where the superscript zero is used to emphasize that these are the
zero-order results. Substituting these results into the terms propor¬
tional to a, we have
TT 2
= U cos (o:ct + 0) H-— + ^aa\ o(l — cos 2a>cf) (1.5.22)
and
(
F0 — ) cos wct + ^—|- -^aaV0( 1 — cos 2wct) (1.5.24)
Finally, from Eq. (1.5.19), if we integrate and apply the initial condi¬
tion on vz, we obtain
which is just the (approximate) equation for the held line passing
through the origin, obtained in Eq. (1.5.12). Thus, one effect of the
curvature is to provide an acceleration which causes the particle to
stay rotating about the same “magnetic held line.”
Of the perturbation terms in vy, the hrst two again show small
1.5 MOTION IN A NONUNIFORM B FIELD 49
VD aya F|<zlo +
a
a.-G^Yd- IV
01 r.
(1.5.27)
which arises partly from the gradient and partly from the curvature.
Note that this drift is in opposite directions for oppositely charged
particles, because of oic. The factor aay can be expressed in terms of
either the gradient in Bz, VBZ = aB0ax, or in terms of the radius of
curvature, R = — oVa*, since with B0 = B0az we have
B0 X R X B0
aay (1.5.28)
B02 R2B0
To2 Bo X TBZ IV R X Bo
(1.5.29)
VD 2cc Bo2 ^ oic R2B0
Fig. 1.19 Typical orbits for charged particles with a small gradient or curvature
in B.
50 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
2 sin u>ct
1 + OM (iff- (1.5.30)
Z = Z°
( COS U)ct
Wct
This result shows that the position of the particle along z can vary
from Z° = Vit by a factor of order oca, which, by hypothesis, is very
small. However, since these variations in Z fluctuate and depend
markedly on the exact orientation of the initial velocity component
perpendicular to B0, they are not classified as a true drift of the particle
in the sense used in Eq. (1.5.27).
Problem
1.13 Estimate the curvature drift velocity for an electron in the earth’s
magnetic field when R is the earth’s radius, B0 = 1 gauss, and I i
is the “thermal velocity” of an electron (kT/m)112, with T = 300°I\.
Thus far we have seen what happens to the motion of a particle when
the magnetic-field intensity changes with position perpendicular to
the “basic” field direction (the z axis, in our problem), i.e., when Bz
is a slowly varying function of x or y. Now consider the case when
Bz is a slowly varying function of 2
or
Solving, we have
aB0r F(z)
Br = (1.5.34)
2 + r
vz = icoca(X°Vy° - Y°vx°)
— —±ucaaVo = — 02 (1.5.37)
Hence
Problem
1.14 Investigate energy conservation for a particle in a converging field
by calculating first-order values for vx and vy and then calculating
±mv2 as a function of time.
SUMMARY
Although our discussion of the motion of a charged particle in non-
uniform B fields has been limited to two rather special problems, these
contain most of the physical aspects of any more general problem, pro¬
vided B does not change markedly over a distance of the order of the
cyclotron radius a. For when B is slowly varying, we can always
approximate B in the neighborhood of the particle by an expansion in
which B is represented by an equation like Eq. (1.5.5) or (1.5.35) (or
a sum of both). This approximation will be good, that is, ax, ay,
and az will all be small compared to unity, for times up to a few cyclo¬
tron rotations; thereafter one can start the problem anew and expand
the field about its value in the neighborhood of the particle’s new
position. Hence, if one is satisfied to employ a computational scheme
of this sort, our solutions can be applied to any nonuniform B field
problem.
M = /A (1.6.1)
t Note that this area A has no connection with the vector potential, A, intro¬
duced earlier. In fact, in the remainder of this text, we shall never again need to
use the vector potential, so that A will almost always denote an area.
1.6 ADIABATIC INVARIANTS; APPLICATIONS; TIME-VARYING FIELDS 53
A*
A B
\ = — . (1.6.2)
and current
I = tv±
q (1.6.3)
‘lira
Adiabatic invariance of \x
dtx . w± w± £ (1.6.5)
di = ^ = -B~B^B
t Note again that B is an externally applied field, not the field produced by the
spiraling charge (which is in the direction of A).
54 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
Aw± (1.6.6)
ht
where Awx is the change in w± during one revolution of the particle and
At = 27t/coc is the time of one revolution. Then, assuming, still, that
the particle’s motion is a closed loop, we have
and
qBvx2 w±B
w, = (1.6.10)
c hr
Using this result in Eq. (1.6.5), we have m = 0.
The proof which we have presented applies to the case when B
is uniform in space but with a time-dependent amplitude. Hence the
motion in the direction of the field is unaffected, and the change in w±
arises from the work done by the E field, which always accompanies a
time-varying B field. In addition, throughout the analysis we neg¬
lected the uniform motion in the direction of B. However, when this
component of the velocity is nonzero, we need only change to a coordi¬
nate system moving with the constant velocity vz. In this moving
coordinate system the particle’s path is in a plane, as assumed in our
proof, and the electric and magnetic fields are unchanged. (This is be¬
cause the corrections involve v2 X B, which is zero.f) The major as¬
sumption in this analysis is that “not much changes” during a single
t See for example, W. K. H. Panofsky and M. Phillips, op. cit., pp. 160-165.
1.6 ADIABATIC INVARIANTS; APPLICATIONS; TIME-VARYING FIELDS 55
dBz/ dz
mvz 9^ — aw± wx (1.6.11)
~BT
dw± ^ w± dBz
(1.6.14)
dt B0 dt
so that
d± = I _ w± dB ~ (16 15)
dt B dt B2 dt
Problem
E = ±B0h(Yax — A"a„)
56 MOTION OF A CHARGED PARTICLE IN ELECTROMAGNETIC FIELDS
= ±much(Yvx — Xvy)
(c) Next, show that for X, Y, vx, and vv given by the zero-order
solutions (valid when h = 0) wJB stays constant.
{d) Find the restrictions on h(t) required by V X H = ecE.
Flux = /B • dS (1.6.16)
Tm2v±2 %rm w±
Flux = Bira2 (1.6.17)
~cflB q2 B
w± _ w±(0)
(1.6.18)
~B ~ 5W
For reflection to occur, however, all the initial kinetic energy must be
transformed into wL \ that is, at reflection, w± = Wtot(0). Therefore, at
reflection we have
wtot(0)
5 = 5(0) (1.6.19)
w±( o)
will reflect one-half the particles, and one of several times B0 will
reflect almost all the particles. However, those particles which
initially are moving in or near the z direction [so that wtot » tt>x(0)]
will escape even if B is very large.
A term commonly used in this situation is the pitch angle a
between v and the 2 axis. If v is the total speed of the particle,
sin a = v±/v, so that, from our definition of m, we have
w± /LiB
sin2 a (1.6.20)
Wtot Wtot
sin2 a
= const (1.6.21)
~B~
which again shows that the pitch angle increases toward 90° (where
reflection occurs), as the particle moves into a region of increasing
field strength.
Jvdl (1-6-22)
where the limits on the integral run over one period of oscillation back
and forth between the mirrors, and where it is assumed that the
separation between the mirrors changes slowly with time. To analyze
this problem we consider the idealized situation shown in Fig. 1.21.
B
-►
V
m
Uniform field
L
—-2
Moving Stationary
magnetic magnetic
mirror mirror
Al — 21 i>s, | L (1.6.23)
with vz the drift speed of the particle in the long region where B is
uniform. The rate of change of AL is clearly
(1.6.24)
with A11;*,| the change in \vz\ in one reflection from the moving mirror
and At = 2L/|v*| the period of oscillation between the mirrors. To
find A|v*|, consider the transformation to a coordinate system moving
with velocity vm. Denote this moving coordinate system with a
prime, so that
vz = v' + vm (1.6.26)
(1.6.27)
or
— —t-'zi + 2 vm (1.6.28)
d\vz\ ^ vm\vz\
(1.6.30)
dt — L
and A L is constant.
The approximate invariance of AL permits one to compute very
simply the energy of a particle trapped between the two mirrors, since
ttiAl2
wz = ±mvz2 (1.6.32)
~8L~2~
wl
M = = const (1.6.33)
MAGNETOIONIC
THEORY
chapter two
u = — vu (2.1.2)
2.1 THE LANGEVIN EQUATION 63
Hence, if there are no external forces acting, the collision term causes
the average electron velocity to approach zero at a rate governed by
the collision frequency between the electrons and the heavy particles.
Note that this does not mean that the velocities of the individual elec¬
trons approach zero due to collisions with the heavy particles but only
that the velocities become completely random, so that the average
velocity is zero.
In addition to Eq. (2.1.1) one might expect to use an analogous
equation for the average ion velocity of the form
However, since m,-, the ion mass, is typically about 104 times larger
than the electron mass, the acceleration of the ions by electromagnetic
fields is generally, very small. Therefore, while there are some prob¬
lems in which the motion of the heavy ions is important, we shall
neglect these for now and assume that the average velocity of all the
heavy particles stays zero.
J = — n0eu (2.1.5)
V x E = —B V X H = D — n0eu (2.1.6)
We now assume that, aside from the effects due to this electron current,
the plasma medium acts like a vacuum with D = eoE and B = MoH,
so that in terms of just E and B we have
V X E = -B V X B = Mo n0eu (2.1.7)
their present form are not yet quite tractable because of the nonlinear
term u X B (involving the product of two variables), which enters
Eq. (2.1.1).
This difficulty can generally be eliminated by the following
technique. First, divide B into two parts
V x E = -b (2.1.9)
Now, for traveling plane waves we shall show in Sec. 2.4 that E, b,
and n are all proportional to exp (tk • r — iut). Hence, the opera¬
tions V x and d/dt become
and
d
(2.1.11)
= ~lW
k X E
l> = - (2.1.12)
CO
where the last term on the right is the nonlinear one. Note, however,
that the magnitude of this nonlinear term is always less than \ukE/u\,
whereas the first term on the right has magnitude \E\. Hence, the
nonlinear term can be neglected, provided \uk/co| is much less than unity
2.2 ELECTRON CONDUCTIVITY AND MOBILITY 65
or
CO
u\ « (2.1.14)
k
0 = — — (E + u X B0) — m
m (2.2.1)
V X E = 0 Vxb = MoJ
E + u X Bo = 0 (2.2.3)
Hence, Eq. (2.2.3) is often called Ohm's law for a plasma with infinite
conductivity. Note that if we take the cross product of each term
66 MAGNETOIONIC THEORY
E X Bo
>h (2.2.4)
B02
The generalized Ohm’s law is not yet in a form where one can easily
calculate J for a given E field. Therefore, in Eq. (2.2.2) we choose our
coordinates so that Bc lies in the 2 direction, and in addition we replace
u by — J/en0, to have
eB 0
J —■ (JoE H-J X a3 LCr (2.2.5)
v m
w,
fj x V to Ex
CO
~Jx + Jy T0B y (2.2.6)
V
J. Tq Ez
From the last of these equations it is evident that a0 governs the flow
of current in the direction of B0, while from the first two equations
we find
with
cror2 <T0C0CV
(Tk (2.2.8)
o>c2 + r2 C0C2 + V2
function of ctjv. It can be seen that for wc/v > 1 the perpendicular
conductivity falls off very rapidly, so that very little current can be
driven directly across the lines of a magnetic field. The Hall con¬
ductivity also falls off with increasing wc/v but at a much slower rate.
Electron mobility
The electron mobility /ie is just the ratio of the average electron velocity
to the electric-field strength in a steady-state problem with no B field.
Thus from the first equation in (2.2.1) we have
u e
He = (2.2.9)
E mv
|u X b| « \E\ (2.2.10)
J = a,Eaz (2.2.11)
K = (2.2.13)
Jxb H()Oo2E2r
u X b = (2.2.14)
n0e 2 n0e
2 n0e
E « (2.2.15)
ju0cr02r
f Note that the term in (V X bh involving d/d<f> is zero because of the sym¬
metry in the 4> direction.
2.3 CONDUCTIVITY TENSOR AND DIELECTRIC TENSOR 69
v = 6.1 X IQ"9
^3oo yi*
n0 (2.2.16)
and
(2.2.17)
300V n0 1
E « 1.2 X 106 volts/m (2.2.18)
T ) 1020 r
Dc conductivity tensor
Jz <T 0 Ez
f This is a maximum value for cr0 (and hence a conservative estimate for E)
since, more generally, v also includes a term for electron-neutral collisions.
70 MAGNETOIONIC THEORY
a • E (2.3.2)
with
O’x <rH 0
ct = <7 j_ 0 (2.3.3)
— <JH
0 0 (T o
Ac conductivity tensor
Equation (2.3.1) is the solution to the linearized Langevin equation
when u = 0. If we instead assume that there is a sinusoidal dis¬
turbance, so that u and E vary with time as e~wt, we have, from
Eq. (2.1.1),
u — icon = —- (E + u X Bo) yu
m
or
Problems
2.1 Consider the following equations for average electron and ion
velocities when B0 = 0:
mu = —eE — mv( u — U)
MV = eE — mv(V — u) M m
2.3 CONDUCTIVITY TENSOR AND DIELECTRIC TENSOR 71
, _ n0e2E 1 + (m/M)
mv 1 + (m/M) — i( w/v)
Thus, for m <<C M, the effect of the ions is truly negligible here for
all co. (This is not the case if B0 ^ 0.)
2.2 Prove that if a solid has equal numbers of electrons and holes
and if mu = me and vh = ve, then the conductivity tensor has
the form
0 0
u = <Tj. 0
0 (To
Dielectric tensor
V X b = — i'(coMoeo)E + MoJ
V x b = — fco/io£ * E (2.3.7)
(2.3.8)
The tensor e which enters here is called the dielectric tensor for the
plasma.
The use of the dielectric tensor in a sense represents a different
approach to treating the plasma from that used till now. Up to this
point we have represented the plasma as a collection of particles moving
about in an otherwise evacuated medium. Thus, in Maxwell’s equa-
tions, we set
as would be the case for a vacuum, and the effect of the plasma shows
up through the currents arising from the motion of the charged par¬
ticles. If, instead, we adopt the dielectric tensor given in Eq. (2.3.8),
we essentially set
D = e • E (2-3.10)
Plane waves
We now want to use the Langevin equation for a plasma (plus Max¬
well’s equations) to study the propagation of plane waves in an
unbounded plasma. Before beginning the analysis, however, it is
useful to review briefly some of the fundamental features of waves.
A typical plane wave traveling in the z direction has the form
Fig. 2.3 Plots of the amplitude of a plane wave as a function of space and time.
2.4 WAVE PROPAGATION WITH Bo = 0 73
that the wavelength X and the period T of the oscillations are given by
2ir 1
X (2.4.2)
T CO
/
with / being the frequency in cycles per second. For any wave there
is a functional relationship between k and co which depends on the
medium and which is known as a dispersion relation. Hence, in Eq.
(2.4.1) we can, alternatively, think of k as being a function of some
applied frequency co or of co being a function of some specific wave
number k.
For a plane wave traveling in some arbitrary direction specified
by the unit vector a
and the phase velocity vph, which is the velocity of propagation d£/dt
of these planes, is found [by differentiation of Eq. (2.4.4)] to be given
by the familiar relation
vph = l (2.4.5)
\p = f A{k)eikz~iut dk (2.4.6)
74 MAGNETOIONIC THEORY
where A(k) is the amplitude (per unit wave number) of the waves in the
disturbance with wave number k. If the range of values for k is small
and centered about some specific value k0, we can let
k — ko k and CO = COo 4*
(2.4.7)
where
(2.4.9)
Equation (2.4.8) has the typical form for a wave packet, such as
the one shown in Fig. 2.4, where co0 is the carrier frequency and M(z,t)
is the amplitude of the packet. If the total range of wave numbers is
small, so that for all k
1
2
kH «1 (2.4.10)
Fig. 2.4 A typical wave packet. The solid curve is ^{z), and the
dotted one is the amplitude factor M(z,t).
2.4 WAVE PROPAGATION WITH B„ = 0 75
then the underlined term in the exponential (and all subsequent terms
which would involve still higher powers of k) are negligible.f Hence
the planes of constant packet amplitude are defined by
z t = const (2.4.11)
v0 =
(2.4.12)
It should be clear from this derivation that the concept of the group
velocity is not too precise, since for any dispersive medium (where w
is not a linear function of k), t eventually becomes large enough to
violate Ecp (2.4.10).
In addition, some modification of both the phase and group
velocity is required when there is damping. The most common type
of damping is simple exponential decay of a signal in space (for a
traveling wave) or in time (for a standing wave). In the former
case k has an imaginary part, and in the latter case w has one. Thus
for damped traveling waves k = 13 + ia, so that for a wave in the z
direction ^ = Ae~az+iPz~iat, and the phase velocity becomes w//3. The
concept of group velocity can also be extended; however, we shall not
go into that problem here since we never have to calculate the group
velocity of damped waves in this text.
Problem
V X E = icob (2.4.13)
V X b = i —; E — (2.4.14)
c2
II = (E 4 n X Bo) (2.4.15)
m( v — iw)
V X (v X E) —— E — IW/XoTloCU (2.4.16)
cl
E
ii —a • — (2.4.17)
n0e
Results with B0 = v = 0
Because of the complexity of Eq. (2.4.17), there is some advantage in
treating the problem in stages. First, if there is no dc magnetic field
and collisions are neglected (B0 = v = 0), we have, from either
Eq. (2.4.15) or (2.4.17),
le E
n =- (2.4.18)
mco
(a>2 — we2) E
V X (V X E) (2.4.19)
c2
useful identity
1/2
coeXD = 3.90 X KEF1'2 = (2.4.20)
V x = ikax X (2.4.21)
E = Exax + E, (2.4.22)
(2.4.23)
or
and
f ar * ( ax x E,)
The first of these results shows that for a transverse wave (E, ^ 0) the
dispersion relation is
while, for a longitudinal wave (Ex ^ 0), we have, from Eq. (2.4.25),
c
vph = - = [1 - (co,2AP)]1/2 " > (2.4.28)
^oo co < co.
A plot of phase velocity vs. frequency is also shown in Fig. 2.6, where
it can be seen that vvh is infinite for co < co, and then decreases rapidly
toward c as co > ». In the “propagation band” (co > co,) we also
note that if Eq. (2.4.26) is differentiated with respect to k, one finds
C2 = VphV0 (2.4.29)
Fig. 2.6 Plot of phase velocity, group velocity, and attenuation factor for trans¬
verse waves with B0 = v = 0.
2.4 WAVE PROPAGATION WITH B0 = 0 79
so that the group velocity is given by the dotted curve in Fig. 2.6.
Note th^t although vPh is always greater than c, the velocity vg that a
signal would have is less than c, as required by the special law of
relativity.
For the longitudinal waves, Eq. (2.4.27) shows only that oscilla¬
tions (so-called plasma oscillations) can occur just at the plasma fre¬
quency. In Sec. 3.4 when we reexamine the wave-propagation prob¬
lem using a somewhat more complicated model for the plasma, we
shall see that this result is the limit one obtains when the electron
temperature goes to zero. Hence this is often called the cold-plasma
result. At this point, however, we can say nothing about the wave
number or phase velocity for the longitudinal plasma waves.
Problems
(Re A • Re B) = JL
2 Re (A • B*)
and
(Re A x Re B) = \ Re (A x B*)
Note that Re denotes “the real part of” and for any function /(f)
<S> = 1 Re (E x H*)
we have
(2.4.30)
or
k F
Ht = —
uyo
1 ki + k0 r ko + ki
T = EtE*
with
Et = l\ 0—ikl
and
(1 — n)(l — n l)eikl -f- (1 + n)(l + n l)e
kc
n = —
w
EtE* ~
cosh2 al + ^(17 — t] 1)2 sinh2 al
ie E
(2.5.1)
mu>[ 1 + i(v/u>)]
We2 |
V X (V X E) E (2.5.2)
c2[l +i(v/ w)]j
we 2
kV = CO2 - (2.5.3)
1 + i(4«)
2.5 THE EFFECT OF COLLISIONS 83
co.
(2.5.4)
1 + i(v/a)
or
iv + \/ 4coe2 — v2
CO = (2.5.6)
Ex = const (2.5.7)
Problem
2.8 Plot cor and a as a function of v/ue.
k2 = A + iB (2-5.8)
with A and B real. First, if we separate k into its real and imaginary
parts, we have
k = (3 + ia (2.5.9)
or
A = P2 - a 2 B = 2(3a (2.5.11)
84 MAGNETOIONIC THEORY
CO 2 = A + iB (2.5.14)
where we let co = cor — ia, then for B negative the wave is damped,
while for B positive it grows.
C0e2 fcO<,2(,/co)
k2c2 = co2 (2.5.15)
1 + (,/co)2 + 1 + (,/co)2
f Note that we assume here that the phase and group velocities are in the same
direction (as they are in all the dispersion equations we shall encounter here).
2.5 THE EFFECT OF COLLISIONS 85
require the use of Eq. (2.5.13) and quickly lead to complicated alge¬
braic expressions, which we shall not bother to go through here. The
results, however, are rather surprising and well worth describing.
First, for w < coa (where «„2 = 3wc2/4), the damping factor a is a maxi¬
mum when v = 0 and decreases continuously with increasing v. (This
is, perhaps, not too surprising since is below the plasma frequency;
hence for o> < con, a is already rather large when v = 0.) Next, for
w > a plot of a as a function of v has the form shown in Fig. 2.8,
where it can be seen that the damping factor first increases with
increasing v, then reaches a maximum, and thereafter falls off toward
zero as v —* <x>. The maximum for a is found to be when
(2.5.16)
so that the plasma behaves like a vacuum, in this limit. The reason
for this behavior is relatively obvious if one examines the ac con¬
ductivity for the plasma derived in (Eq. 2.3.5):
r nne2
a0 (2.5.18)
m{v — iu)
For large v, this approaches zero, so that no current flows in the plasma,
and the power absorbed (which is the product of the real parts of J
and E) goes to zero.
Problems
2.9 Verify that for w2 = A + iB the wave is damped when B is nega¬
tive and growing when B is positive.
2.10 Show that for w » coe the damping factor for transverse waves
is given by
~ a*Z 7 = -
“ — 2o>c(l + Z2) co
and
Although there are a number of ways to find the dispersion relations for
plane waves from these equations, probably the one which involves
the least algebra is to assume again, as in Sec. 2.4, that E and u are
both proportional to elkx~lwt, then use Eq. (2.6.1) to find E in terms ol u,
and finally use Eq. (2.6.2) to obtain an equation involving just u.
With this form for E, Eq. (2.6.1) becomes
CO2
— k2a* X (a, X E)-y E = — iwju0n0eu (2.6.3)
0
a* X (ax X E) = — E( (2.6.4)
2.6 RESONANCES AND REFLECTION POINTS WITH Bo ^ 0 87
CO
in0eux
Ex (2.6.6)
C0€0
and
in0eut
(2.6.7)
coe0(l — n2)
where the first two terms on the right are perpendicular to a* and the
last is parallel to ax. Hence if we substitute Eqs. (2.6.6) to (2.6.8) into
Eq. (2.6.2) and remember that (n0e2/me0)1/2 is the plasma frequency
cog, we obtain from the x component of each term
lew , -r, ,
ux( CtT T" iwv) = - (u( X Bor) (2.6.9)
m
i te w
u, + IWV I =- (ux X Bor T u( X Box) (2.6.10)
m
Thus we now have three scalar equations involving just the three
components of u. To obtain a dispersion equation, we choose a set of
rectangular coordinates as indicated in Fig. 2.9, where ax is the previ¬
ously assumed direction of wave propagation and az is chosen per¬
pendicular to the plane formed by ax and Bn. If we let y be the angle
between a* and B0, we have
Writing out Eqs. (2.6.9) and (2.6.10) explicitly and dividing through
by w2, we obtain an equation of the form
A • u = 0 (2.6.12)
where
1_
— iYsin y
b
0
1
A = 0 U — iY cos 7 (2.6.13)
iY sin 7 — iY cos 7 U — <3?
and
k2c2
U = 1 +i- X - n2 =
CO CO2 or
(2.6.14)
eB0 ^ = -
(J) x
Y =
TOO) 1 — n2
det A = 0 (2.6.15)
F2 sin2 1/2
7 F4 sin4 7
<£=£/ — 2 + F2 cos2 7 (2.6.18)
2 (17 - X) - 4(U - X)
with <t> given by Eq. (2.6.18). This last result is the Appleton-Hartree
dispersion relation, which is used with considerable success to study
radio-wave propagation in the ionosphere, taking account of the earth’s
magnetic field. For the most part, the results predicted by the equa¬
tion remain unchanged even when more sophisticated models of a
plasma are used, provided the wave frequency is large compared to
the ion cyclotron frequency. For lower frequencies, the ion motion
becomes important, and, as shown in Chap. 3, there are quite different
types of waves which are not predicted by Eq. (2.6.19).
Resonances
E = in°eUt _ (2.6.20)
‘ cueoU - n2)
00 (2.6.21)
90 MAGNETOIONIC THEORY
U sin2 y\
U 2 = Y2 ^cos2 7 + (2.6.22)
U - X)
or
2\2 1/2
We2 + C0c2 We
+ we2wc2 sin2 7 (2.6.25)
Fig. 2.10 Frequencies which satisfy the necessary condition for a resonance as
a function of the angle 7 between B0 and the direction of wave propagation.
The curves on the left are for we > |oc|, the curves on the right for coe < jcoJ.
2.6 RESONANCES AND REFLECTION POINTS WITH B„ ^0 91
Problem
Neglect collisions, and start the particle at the origin with zero
velocity. Note that this is a nonphysical problem since the fields
violate Maxwell’s equations.
Reflection points
(U - X)2 - F2 = 0 (2.6.26)
(1 — X) = + Y
or
so that
(The minus sign in front of the square root is not considered, as this
leads to a negative frequency. Also, we use |wc| since a>c is negative.)
Thus, reflection points for any angle of propagation occur at two fre¬
quencies, wqi and co02, which lie one on each side of we and which are
separated by ay rad/sec.
92 MAGNETOIONIC THEORY
Problems
2.12 Prove that co0i < coe < \/coc2 + coe2 < coo2, so that the reflection
points lie, respectively, below coe and above the hybrid resonance
frequency.
2.13 Derive an equation analogous to Eq. (2.6.18) for a system with
equal number densities of electrons and holes and with m* = m
and vh = v. What are the reflection points and resonances for
such a plasma?
Propagation parallel to B0
To continue our discussion of waves in the plasma we now examine the
Appleton-Hartree equation for some special cases. First, for wave
propagation in the direction of B0, 7 = 0, so that from Eqs. (2.6.18)
and (2.6.19)
4> = U + Y U = 1 + i- Y = — (2.7.1)
w CO
2 .,
2 _ 2 _ _—_ (2.7.2)
1 + i(v/u>) + |o)c|/«
If we neglect collisions, we can easily find for the root with the plus sign
that there is no resonance but there is a reflection point (k = 0) when
Other than this, however, the dispersion relation for this wave is very
similar to the simple relation when Bo = 0, as can be seen in the plot of
phase velocity vs. frequency shown in the solid curve of Fig. 2.11.
2.7 WAVE PROPAGATION PARALLEL OR PERPENDICULAR TO Bo 93
When we use the minus sign in Eqs. (2.7.1) and (2.7.2), there is a
resonance (/c = °o) for w = |coc|, as expected from Fig. 2.10, and a
reflection point (k = 0) at w02 = woi + |coc|. Furthermore, k2 is posi¬
tive both for co > co02 and for co < |coc|, so that the phase velocity is
finite for two ranges of frecpiency, as shown in the dotted curve of
Fig. 2.11. (In this discussion and in the figure we continue to neglect
collisions. Otherwise, as we have noted, there are no resonances or
reflection points, and some propagation can occur at all frequencies.)
Note that there is no resonance at coe for either of the waves,
although one is predicted there in Fig. 2.10. This is apparently due
to the fact that in deriving the resonance condition we multiplied by
1 — X, which is zero at w = we. At all other directions of propaga¬
tion, however, there is no difficulty with the resonance conditions.
—iuy + uz = 0 (2.7.5)
or
uy = ±iuz (2.7.6)
Fig. 2.11 Plot of phase velocity vs. frequency for waves traveling in the
same direction as B0, the applied magnetic field.
94 MAGNETOIONIC THEORY
Ul = Uoei^-iat-ax (2.7.7)
uy = iuoe*13*-™1-011 (2.7.8)
t Alternatively, for an observer looking into the lines of B0, u rotates clockwise
(like a -positive charge, as discussed in Sec. 1.4).
2.7 WAVE PROPAGATION PARALLEL OR PERPENDICULAR TO Bo 95
Propagation perpendicular to B0
For wave propagation perpendicular to B0, 7 = 90°, so that, from
Eqs. (2.6.18) and (2.6.19), we have
*- u + mr^x)(-1 ± l) (2-7'1J)
and
k2c2 = “2 ~ 11 , ,•/ / n
+ l(jyco)
, r[COc 2/( — 1i^n/n/
+
2
± 1)/2(C02
1 ,•-iyj
+ lOJV — U>e )J
(2-7.i2)
(U - 4>)Uy = 0 (2.7.13)
Hence, we can have a linear polarized wave (uy 9^ 0) for 4> = U, which
is precisely the dispersion relation given in Eq. (2.7.11) using the plus
sign. In other words, the root with the plus sign corresponds to a
wave involving electron velocities solely in the y direction. Since
this is also the direction of B0, the magnetic force term u x B0 is zero
for this wave, and the wave propagates as though J30 were zero.
If we neglect collisions, we easily find for the remaining root (with
the minus sign) in Eq. (2.7.12) that there is the expected resonance at
(coc2 + coe2)1/2 and the reflection points when co = w0i and w02. Hence
the plot of phase velocity vs. frequency has the form shown in the
dotted curve in Fig. 2.13, where it can be seen that there is propaga¬
tion for co > W02 and for u in a band between co0i and (coe2 + coc2)1/2.
For other frequencies the phase velocity is infinite, since k is imaginary.
96 MAGNETOIONIC THEORY
The two special cases we have considered are both unique in that the
resonance at coe is absent at y = 0° and the low-frequency resonance is
absent at 7 = 90°. For all other angles of propagation there are both
high- and low-frequency resonances, so that for the usual case when cce
exceeds |coc| the curve of phase velocity vs. frequency resembles both
the low-frequency portion of Fig. 2.11 (with the resonance shifted to
some value below |«c|) and the high-frequency portion of Fig. 2.13
[with the resonance shifted to some point between aje and (coe2 + wc2)1/2].
This is shown in the top part of Fig. 2.14. Probably the most impor¬
tant points to note are (1) that no waves can propagate at frequencies
between the lower resonance frequency and the lower reflection point;
(2) that the reflection points are independent of 7; (3) that the lower
resonance frequency decreases with angle from |o>c| when k is parallel
to B0 down to zero when k is perpendicular to B0; and (4) that the
higher resonant frequency (which is missing only when k is exactly
parallel to B0) increases with angle but always lies between and
(cOe2 + CO,2)1/2.
For the case when is less than |wc|, which is unusual in a plasma,
we have noted earlier that the roles of we and |coc| are interchanged in
the resonances, so that the phase-velocity plot has the form shown in
the lower part of Fig. 2.14. The major difference in this case is that
since w0i is always less than we, there is no longer any frequency band
where waves cannot propagate.
Fig. 2.13 Plot of phase velocity vs. frequency for waves traveling in a
direction perpendicular to B0, the applied magnetic field.
2.7 WAVE PROPAGATION PARALLEL OR PERPENDICULAR TO B0 97
In this discussion we have seen that for the usual case when we is
much larger than |wc|, wave propagation at frequencies below |coc| is
restricted to a cone of angles about the direction of B0. The maximum
angle at which propagation can occur is given by the lower resonance
curve in Fig. 2.10. For uc2 « ue2 one easily finds [from Eq. (2.6.25)]
that
Thus, for a) = (\/3/2)|coc|, Ymax is 30°, and for « = |-|coc|, ymax is 60°.
When « is well below |w0|, the group velocity dw/dk is easily found to
increase with increasing frequency (see Prob. 2.14). Hence if a pulse
Fig. £.14 Plot of phase velocity vs. frequency for waves traveling in
an arbitrary direction relative to B0, the applied magnetic field.
98 MAGNETOIONIC THEORY
Problems
2.14 Prove that, for the whistler mode with v = 0, when co is much
less than either the plasma or cyclotron frequency, we have
Vph = C V«lwc| COS y/ue and v„ = 2c y/co|coc| cos y/ue, so that vPh
Fig. 2.15 Plot of 0 versus co for waves traveling in the same direc¬
tion as B0, the applied magnetic field.
From this equation one can easily obtain the plot of /3 versus co shown
in Fig. 2.15.
On the j3 versus co plot, the resonances are where f /? —> °° , and the
reflection points are where f3 = 0. From the plot one can also calcu¬
late both vph = co/13 and va = dco/d/3. In fact, by comparing the
curves with the line /3 = co/c, we note that:
1. If /3 is less than co/c, we have vph > c, while for (3 > co/c, we have
Vph ^ C.
2. If the slope d(3/dco is greater than the slope of the co/c curve (as is
the case in the figure), we have vg < c, as required by special
relativity.
f An exception to this rule is when co = 0, since in this limit vpt, can be zero
even when /3 is finite or zero.
100 MAGNETOIONIC THEORY
Fig. 2.16 The CM A diagram for the Appleton-Hartree equation. The solid
lines are the loci of the reflection -points, where vvh = «, and the dotted lines
are the loci of the resonances, where vph = 0.
The CM A diagram
and
Hence the loci of the reflection points lie on the solid curves shown in
Fig. 2.16, which are defined by these two equations. In addition the
1 - X X - 1
F2 = (2.8.4)
1 — X cos2 y A^ cos2 7 — 1
For 7 = 0 this is the straight line l"2 = 1, and for 7 = 90° it is the
straight line F2 = 1 — X. However, for 7 ^ 0 or 90° one can easily
verify that Y2 has two branches, for which both X and Y2 are positive.
(Negative values of Y2 or A" are, of course, not physically meaningful.)
I11 the first branch, which is the high-frequency resonance, Y2 decreases
monotonically from 1 when A" = 0 to 0 when X = 1. In the second
branch (the low-frequency resonance) F decreases from 00 when
A" = sec2 7 to sec2 7 when A —> °o. Hence, from Eq. (2.8.4) one can
easily construct the resonance curves for 7 = 0, 30, 60, and 90°. All
these (except the low-frequency resonances at 60°, where X and F2
are both always greater than 4) are shown as dotted curves in Fig. 2.16.
The two reflection-point curves and the resonance curves for 0 and
90° divide the XY2 plane into eight regions. With some additional
analysis one can make a polar plot of the phase velocity as a function
of angle 7 in each of these regions, so that plase-velocity information
is also included in the CMA diagram.
In summary, there are any number of ways of visualizing the
information contained in a dispersion relation like the Appleton-
Hartree equation. Each has its advantages, and it is probably usefid
to be familiar with a number of these different types of diagrams.
However, in the next chapter, when we again analyze wave propaga¬
tion in a plasma using a more detailed model of a plasma, we shall for
the sake of brevity use only the simple diagrams of phase velocity vs.
frequency.
chapter three
CONTINUUM
EQUATIONS
FOR A PLASMA
chapter three
Macroscopic methods
Mass-conservation equation
The total mass contained in the volume is J'p dV. Hence, for the
volume element
p + V • pu = 0 (3.1.4)
Momentum-conservation law
DF dF , dF , dF . dF dF
+ u • VF (3.1.6)
TvT = ITT + Ux “I" a- uv “r XT Uz dt
Dt dt dx dy dz
f = — (E + u X B) (3.1.8)
m
To find the net force due to the pressure suppose the mass, at a time t,
is contained in the rectangular volume element shown in Fig. 3.2.
Then the net force acting in the x direction is clearly (recall pressure is
force per unit area)
V
(
(z -
5x
y* y>z)\ - p[■x
/ , Sx \
+ ~2> y>z) by bz (3.1.9)
P(R) - - p(R)
dp M dp
bx by bz (3.1.10)
dx p dx
3.1 DERIVATION OF CONTINUUM EQUATIONS 107
where in the last step we have noted that the mass M equals the
density p times the volume Sx by bz. From Eqs. (3.1.7), (3.1.8), and
(3.1.10) we have, for our simple gas of identical particles,
(3.1.11)
~ = 2- (E + u X B) - - Tp (3.1.12)
Dt m p
Many other forces might be added on the right side of this equation:
the gravitational force on the mass element, the interaction force due
to collisions of these particles with particles of other species (which
we add in the next section), and various viscous forces due to stresses
acting on the mass element. However, in many problems the electro¬
magnetic force and the pressure-gradient force, which we have included
here, are of predominant importance, and one can safely neglect all
the other force terms.
Problem
3.1 Derive the mass-conservation equation by letting F — p and not¬
ing that M = p bV is constant. [You must actually calculate
D(bV)/Dt from a figure like Fig. 3.2.]
Although one can derive a general energy equation using the same
techniques already used to derive the mass and momentum equat ions
we shall for now restrict attention to two special cases, the isothermal
and the adiabatic energy equations. As will be shown in the chapters
on kinetic theory, the scalar pressure p satisfies the condition
p = nuT (3.1.13)
where (p/p)112, which enters here, has the dimensions of velocity and is
called the isothermal sound speed. More generally, when the tem¬
perature inside the plasma is not fixed, wTe shall merely assume (as
is often done in hydrodynamics) that any changes in pressure or volume
satisfy the familiar adiabatic lawf (commonly used in thermodynamics)
or
dp = — clp = U2 dp (3.1.17)
P
f The conditions which are required to obtain this adiabatic law are investigated
in Sec. 4.2.
3.2 TREATMENT OF PLASMA AS A MIXTURE 109
with l = (yp/p)1/2 the adiabatic sound speed for the fluid. From
Eqs. (3.1.14) and (3.1.17) we sec that the isothermal or adiabatic
assumptions mean that a change in pressure is accompanied by an
instantaneous, proportional change in the density. For many problems
this is a good approximation; however, as one might expect, in reality,
if the pressure change in a gas is very rapid, there is a time lag before
the density of the gas can respond. Hence for problems involving a
rapidly fluctuating pressure, one must replace these simple laws with
a more general energy equation.
The continuum equations we have developed thus far have the form,
for a gas of particles of type a,
Pa + V • (Pau„) = 0 (3.2.1)
+ <3-2-2>
and
= UJ VPa (3.2.3)
with Ua, depending on the problem at hand, either the adiabatic or the
isothermal sound speed. As we noted in our derivation of the momen¬
tum equation, there are other forces which may act on the mass element
of particles of type a. In particular, if we have a gas mixture, there
should be a force term acting on the mass element of particles of type a
due to collisions between particles of type a and particles of other
species. A simple way to take some account of this force is to assume
that the force per unit mass on the a-particle gas due to collisions with
particles of some other type (3 is proportional to the velocity difference
between the two gases, so that fap — — rap(ua — up), and we have
^
l~)t
^
Wla
(E + ua x B) - ^-
Pa
y Vap(ua - Up) (3.2.4)
(3-2.6)
PaVa!3 — PWa V '
(3.2.7)
(3.2.8)
where we have used the fact that for a uniform gas, from Eq. (3.2.1),
pa = 0. Using Eq. (3.2.7) to eliminate ua, we have
(3.2.9)
Now, for any function which is summed over two indices, the result is
unchanged if we interchange the indices. Hence, for any function
fa& we have
(3.2.10)
or
(3.2.11)
3.2 TREATMENT OF PLASMA AS A MIXTURE 111
Since each term on the right is positive, W is negative, and the kinetic
energy decreases until all the individual fluid velocities are equal.
To find the equilibrium value for all the individual fluid velocities,
consider the total fluid velocity
(3.2.13)
Thus, setting ua( °o) = u = u(0), we have for the equilibrium velocity,
Problem
3.2 Solve Eq. (3.2.7) for ua(t) when there is a two-fluid mixture and
when there is a three-fluid mixture. (The three-fluid mixture
can best be solved by Laplace transforms.)
f As usual, the square of a vector is defined as the dot product of the vector
with itself.
112 CONTINUUM EQUATIONS FOR A PLASMA
Problems
3.3 Show that if the energy equation for a homogeneous mixture with
no external forces is
2rnavaji
T„
-1 m„ + m,3
('Ta - Tf,) - ^ |u« - U/3|2
To see how the present set of equations compares with those used in
Chap. 2 in our discussion of magnetoionic theory, note that if we
assume that the flow velocities of the heavy particles are all zero, we
have, for the electron “gas,” from Eqs. (3.2.1) and (3.2.4),
p + V • (pn) = 0 (3.2.16)
and
u r n • Vu — — (E + u X B) — - Vp — ; u (3.2.17)
m p
3.3 DIFFUSION 113
where all quantities refer to the electrons and v = 2pea is the total
electron collision frequency for momentum transfer with all types of
other particles. As noted earlier, Eq. (3.2.16) is entirely equivalent
to the charge-conservation law, since if we multiply by e/m and note
that p = mne, we have
Pc + V • J = 0 (3.2.18)
Ue2 VP
(3.2.19)
P
3.3 DIFFUSION
n + w0V • u = 0 (3.3.2)
kT
u + u • Vu Vn — p u (3.3.3)
mn0
114 CONTINUUM EQUATIONS FOR A PLASMA
kT „ . (3.3.4)
—n — V2n + vn
m
or
(3.3.5)
n = De V2n-n
V
where
(3.3.6)
De = —
mv
n n n
n ~ - De V2n — • (3.3.7)
r V t2v
Comparing the first and last of these terms, we note that if the
average number of collisions for each electron during the time r is
large, that is, vt 1, then the last term [in Eq. (3.3.5) or (3.3.7)] is
small compared to the first, and Eq. (3.3.5) reduces to the diffusion
equation
n = De V2n (3.3.8)
f Note that, in the usual fashion, we also assume that all derivatives of n and
u are infinitesimal.
3.3 DIFFUSION 115
n Den
- - (3.3.9)
rd L2
or
L2
td (3.3.10)
De
Problem
(d) Find n(x,t) when n0(x') = e-x'2/*°2 and note that td = x02/D
is the characteristic time for the diffusion to smooth out n.
kT
L2 » mv2 (3.3.11)
116 CONTINUUM EQUATIONS FOR A PLASMA
V • E = (3.3.12)
€o
(3.3.13)
to
eE e2nL
/e = -
(3.3.14)
m me o
Similarly, from Eq. (3.3.3), the diffusion force per unit mass is of order
kT kT n
fo = — Tn ~ (3.3.15)
mn0 mn0L
Hence, the neglect of the electric-field term is justified only iorfE «/o,
or
xTe o
L2 « = Ad2 (3.3.16)
n0e2
(kT/m )1/2
Ad — (3.3.17)
ue
From the plot of Debye length in Fig. 1.1 we have seen that \D
is typically very small. Hence Eq. (3.3.16) is rarely satisfied, and for
most problems the electric-field term cannot be neglected.! Therefore
we shall next reexamine the diffusion problem, taking account of the
motion of both the ions and electrons and retaining the E field.
Ambipolar diffusion
with n'a and the fluid velocities ua all assumed to be of very small
amplitude. With this assumption, we have for each species the mass
conservation equation
(3.3.20)
where /3 takes on the values e, i, and n for electrons, ions, and neutral
molecules. Taking the divergence of this equation and using Eq.
(3.3.19), we obtain (after multiplying through by n0 and assuming
that the neutral velocity u, is zero),
\r>2coei
« L2 « \D2
V
When a)e is larger than v (as is the case for most plasmas), it is clear that the
assumptions used in the theory of free electron diffusion cannot both be met.
118 CONTINUUM EQUATIONS FOR A PLASMA
There are two equations of this form, one each for the electrons and
the ions. Therefore, if we replace V • E by the divergence condition
we obtain two coupled equations for the two variables ft' and n'e.
However, these coupled equations are generally too complicated to
treat in detail except by numerical means unless we make further
assumptions. To start we recall from Eq. (3.3.7) that the term na
can be neglected compared to h'avan when the characteristic time for
diffusion r„ satisfies the condition vanra » 1, so that the average electron
or ion has many collisions with neutral molecules during the time of
diffusion. With this assumption we have, from Eq. (3.3.21),
or
(3.3.27)
with
K(Te + Ti)
(3.3.28)
nie^en I nil V\n
so that
meven ^ /meTA1/2
(3.3.30)
TYliVm \7Yl%T{ J
which is usually very small, since we/mt- is of order 10-4. With this
approximation we have
(3.3.31)
Wli Vin 0 + r) - D< 0 + y)
E ^ Lejn'j - w')
V • E = — or (3.3.32)
eo Co
U ^ LKK ~ K) (3.3.33)
Jd \D2(a)n'a
where A.o(a) = V^T „£0/n0e2 is the Debye length using the temperature
for species a. Since L2 is generally much larger than Ad2, it follows
that the electric-field force (which tends to equalize n' and n’e) becomes
very strong whenever there is any significant deviation from charge
neutrality.
120 CONTINUUM EQUATIONS FOR A PLASMA
Problem
3.6 Use Eq. (3.3.23) or (3.3.24) to demonstrate that when the ambi-
polar diffusion equation is used for n'e or n(, to a good approximation,
K - K XD2
K ~ u
so that, for \D2 « L2, the analysis is at least consistent.
r = -DeVn + — T x a2 (3.3.36)
V
I = - D • Vn (3.3.37)
3.3 DIFFUSION 121
Dl 0
D = -Dh Dx 0 (3.3.38)
0 0 De
and
Dy
D .L
y+ f2
(3.3.39)
DeVCOc
Dh
COc2 + F2
n -f v • r = 0 (3.3.40)
n = V • (D • Vn)
(3.3.41)
Since Dx (like <rx shown in Fig. 2.1) falls off rapidly with increasing
values of |o>c|/f, this result indicates that the diffusion of particles
across the field lines is greatly reduced when |coc| is much larger than v.
Experimentally, it is found that while the diffusion is reduced by the
field, the reduction is not always so great as expected from this simple
theory (where D± is proportional to B0~2 for |wc| » f), and various
instabilities have been proposed to explain this anomalous diffusion
in the experiments.
Problems
11 ,= U <K ^ (3.4.1)
icot
B = Bo + beikx~iut
and
the plasma and u, is assumed zero, we still have [see Eqs. (2.6.6) and
(2.6.7)]
in0eux
Ex (3.4.4)
coe0
in0eut
E, (3.4.5)
coe0(l — n2)
Substituting these expressions for E and Eq. (3.4.2) for n into Eq.
(3.4.3) and multiplying through by ico, we obtain
ie<o
Uj(u2 — W2 + iwv — k2Ue2) = — (u, X B„,) (3.4.6)
m
2
ieu
U, W lWv\ = —
-I- icc (u* x B0, + u, x Bo*) (3.4.7)
1 ni m
From these two cases and a knowledge of the reflection and resonance
points, it is again possible to infer the dispersion curves for waves at
an arbitrary orientation of B0.
Propagation parallel to Bo
For our first example we set B0 = Boax, so that B0, = 0, and Eqs.
(3.4.6) and (3.4.7) become
Ue e_Bo
+ iwv iwcwUz Uc = (3.4.8)
Uy
1 — n2 ) m
U)e2
uz — iwwcUy
1 - n2
f Since Uc2 goes to zero when T = 0, these equations are often called the warm-
electron-plasma equations to distinguish them from the cold-plasma equations in
magnetoionic theory.
124 CONTINUUM EQUATIONS FOR A PLASMA
From the first of these we see that, to have longitudinal waves (ux ^ 0),
the dispersion relation is
(Recall that using the Langevin equation, we had found this result
minus the crucial term on the left.) When collisions are neglected,
note that this longitudinal-dispersion relation is entirely analogous
to the simple transverse-wave dispersion equation (when B0 = 0)
except for the change from c2 to Ue2- Hence a phase-velocity plot
(for v = 0) has the familiar form (shown in Fig. 3.3) of a reflection
point at cce followed by a rapid decrease in the phase velocity toward
Ue as co increases beyond a>e.
The remaining two equations in (3.4.8) do not involve Ue and are
precisely the two transverse equations obtained in magnetoionic theory.
Hence, they lead, again, to the l.c.p. and r.c.p. waves described in Sec.
2.7 with phase velocities (for v = 0) as shown in Fig. 2.11 and repeated
here in Fig. 3.3. Thus the effect of the warm-plasma model for this
case is to add a new longitudinal (or acoustic-type) wave (in place
of the simple oscillations at the plasma frequency) to the two trans¬
verse waves we had already encountered.
The reason that the addition of the pressure-gradient term has no
effect on the propagation of the transverse waves is apparent if we recall
that we have used the adiabatic assumption, Vp = Ue2 Vp, so that Vp
is zero if the density is constant. Now, from mass conservation we
see in Eq. (3.4.2) that for small-amplitude waves traveling in the x
Fig. 8.8 Plot of phase velocity vs. frequency for plane waves traveling in the
same direction as B0, the applied magnetic field, when the warm-electron-gas
equations are used.
3.4 WAVE PROPAGATION IN A WARM PLASMA 125
kn0ux
n = - (3.4.10)
co
Problem
Propagation perpendicular to B0
For our second example let B0 = B0a„, so that Eqs. (3.4.6) and
(3.4.7) become
ue2 i • N n
— icococux + Uz • --, + icov ) - 0
1 — n2 )
CO2 = 0 (3.4.13)
— icococ co2 + icov
1 — n‘
126 CONTINUUM EQUATIONS FOR A PLASMA
or
we
k2c2 = co2 (3.4.16)
1 — C0c2/ (co2 — We
Fig. 3.4 Plot of phase velocity vs. frequency for plane waves traveling per¬
pendicular to B0, the applied magnetic field, when the warm-electron-gas equa¬
tions are used. Note that in place of a resonance at (o)c2 + o><.2)1/2, there is a
transition from a basically transverse wave to a basically longitudinal one.
3.4 WAVE PROPAGATION IN A WARM PLASMA 127
or
(3.4.18)
so that
UM ± (t/AV + 4C/e2cWo>e2)1/2
k2 =
2c2Ue2
(3.4.20)
128 CONTINUUM EQUATIONS FOR A PLASMA
The solution with the minus sign is, of course, an evanescent wave,
while the phase velocity for the root with the plus sign is
1/2
CO (oJe2 + Uc2)cUe (3.4.21)
k
Conclusions
Comparing Fig. 3.4 with the analogous cold-plasma result shown in
Fig. 2.13, it is clear that the principal difference lies in the absence of a
resonance at (we2 + wc2)1/2. Instead of dropping to zero, the phase
velocity drops toward the electron sound speed Ue. This same effect
is found for wave propagation at other angles where one of the reso¬
nances is eliminated, and, instead, there is a transition to the acoustic
mode.
The one remaining resonance is easily found (see Prob. 3.10) to
be at the frequency co = |wc| cos y, which is the value for the low-
frequency resonance found for the cold plasma in the limit when
we » |coc| and used in our discussion of the whistler mode. Hence a
typical plot of phase velocity vs. frequency for an arbitrary angle
between 0 and 90° has the form shown in Fig. 3.5.
Fig. 3.5 Plot of phase velocity vs. frequency for plane waves traveling
at some angle y relative to B0 when the warm-electron-gas equations
are used.
3.5 LONGITUDINAL WAVES IN A FULLY IONIZED GAS 129
Problem
na + V • (naua) = 0 (3.5.1)
where a or /3 takes on the value e for electrons and i for ions (with
/3 5^ a), and where Ua is, again, the adiabatic sound speed for particles
of type a. Restricting ourselves again to small-amplitude waves, so
that
ua = uaeikx~iut ua « ^
(3.5.3)
B = Bo + be**1-1"'
, kTlQUax
na = -- (3.5.4)
CO
UJ
— icoua = — (E + ua X B0) iknaax ^at)9(via *•»*) (3.5.5)
ma n0
130 CONTINUUM EQUATIONS FOR A PLASMA
which is like the current in an electron gas except for the change from
ue to ue — Ui. Hence, in place of Eqs. (2.6.6) and (2.6.7) we obtain
from Maxwell’s equations
(3.5.7)
(3.5.8)
Longitudinal waves
With B0 in the x direction, u„ X B0 is perpendicular to ax. Therefore
when we substitute Eqs. (3.5.4) and (3.5.7) into Eq. (3.5.5) we have,
from the x component of the electron and ion equations (after multi¬
plication by ioo),
me
and (3.5.10)
Mi
Thus for a typical ionized gas the ion collision frequency, plasma
frequency, and sound speed are all much smaller than their electron
counterparts. This enables one to simplify much of the algebra
involved in obtaining the dispersion relations.
To have longitudinal waves (uex, UiX 9^ 0) the determinant of the
coefficients in Eq. (3.5.9) must be zero; that is,
777
^2 = a2— (3.5.13)
»!,■
k4Ue2Ui2 - k2
M)
U2(Ue2 + Ufi) - UfiU2 ( 1 + ^
where the two underlined terms are very small and will henceforth be
neglected. Although it is not difficult to write down the two exact
solutions to this quadratic equation in k2, it is more convenient to
seek out separately some approximate solutions at high and low fre¬
quencies. For this purpose we again neglect collisions (so that
fl2,- = co2t-) and first assume that w2 » wt2[l + (Ti/Te)], so that Eq.
(3.5.14) reduces to
or
£W » - ue2| (3.5.17)
which follows from Eq. (3.5.11) for w2 >>> coi2[l + (Ti/Te)]. From
Eq. (3.5.16) we easily see that for high frequencies the two dispersion
132 CONTINUUM EQUATIONS FOR A PLASMA
Fig. 3.6 Plot of phase velocity vs. frequency for the longitudinal waves
in a fully ionized plasma. This curve is for propagation in the direc¬
tion of B0, the applied magnetic field (and also holds when B0 = 0).
relations are
where
f The approximation is only that Eq. (3.5.20) contains an extra term, — k2U^(t>2,
which, however, is small compared to fc2f7p2co,-2 [the second term in Eq. (3.5.19)
when we divide by mc/m,] for co2 « co,-2(l + Ti/T,).
3.5 LONGITUDINAL WAVES IN A FULLY IONIZED GAS 133
of which the first is a wave traveling at the plasma sound, speed Up,
while the second is an evanescent wave (with infinite phase velocity).
The phase velocity for these solutions appears as the low-frequency
portion of Fig. 3.6. To complete the curve one need only solve the
dispersion equation at w2 = Wi2[l + (Ti/Te)] to have
so that, for the plus sign, the phase velocity (after some simple algebra)
is found to be
Hence
Problems
If we return to the equation for u [Eq. (3.5.5)] and take now the
transverseportion, we have, after multiplying by i/w, neglecting
collisions, and setting B0 = B0ax,
co<*2 eB o
<^>a — 9 19 9 ^ a
(3.6.3)
co“ — k2c2 ma co
with a equal to e for electrons and i for ions. From Eq. (3.5.10)
we have
4>i _ Yj _ me
(3.6.4)
Ye mi
Hence, if we multiply Eq. (3.6.2) by 4>e and use this result, we find
Hence, collecting all terms in Eq. (3.6.6), we have [after using the
identity 4\Fe = <EF;, which follows from Eq. (3.6.4)]
Uey _ u.
Or Uey2 = — Uez2 (3.6.10)
u,
3
N
t Note that in the last expression on the right of Eq. (3.6.13) we use either
both upper signs (for the l.c.p. wave) or both lower signs (for the r.c.p. wave).
136 CONTINUUM EQUATIONS FOR A PLASMA
w 2
k2c2 = w2 — -~—
<t>e
where |coce| = eB0/me and uci = eBo/rrii are, again, the electron and
ion cyclotron frequencies. This result is identical to the dispersion
relations obtained in magnetoionic theory [Eq. (2.7.2) with v = 0]
except for the underlined term in the denominator. Hence, for fre¬
quencies large compared to ccci (where the underlined term is negligible
compared to 1) the phase velocities obtained from Eq. (3.6.14) are
identical to those obtained in magnetoionic theory and shown in
Fig. 2.11 (as well as here, in the high-frequency part of Fig. 3.8).
Alfven waves
When we go to the low-frequency limit, i.e., when we set co <3C wci,
the dispersion relation (3.6.14) reduces for both the l.c.p. and r.c.p.
waves to
k2c‘ = w2 (l + (3.6.15)
coe2 _ n0rrii _ c2
(3.6.16)
| coce| C0ci e0B o2 -Va2
Fig. 3.8 Plot of phase velocity vs. frequency for the transverse waves in a
fully ionized plasma. This curve is for propagation in the direction of B0,
the applied magnetic field.
3.6 TRANSVERSE WAVES IN THE DIRECTION OF B, 137
Bo
Va = 1/2 (3.6.17)
(Mo Pi)
For the usual case when Va is much less than c, Eq. (3.6.15) reduces to
so that .both waves travel at the Alfven velocity Va, as shown in the
low-frequency portion of Fig. 3.8.
Resonances
Recall that resonances occur when k goes to infinity [or the denomina¬
tor in Eq. (3.6.14) goes to zero]. Clearly for the r.c.p. wave, where
we use the lower signs in the denominator, the resonance is still
exactly at the electron cyclotron frequency, while for the l.c.p. wave,
where we use the upper signs in Eq. (3.6.14;, there is a new resonance
exactly at the ion cyclotron frequency, as shown in Fig. 3.8. Again
this resonance is riot unexpected, since we have seen that the natural
motion of an ion in the field consists of left-handed spirals at the ion
cyclotron frequency, so that this resonance is entirely analogous to
the one at |uce| for the r.c.p. wave.
The appearance of this new resonance is one of the most important
features of this analysis. In fact the high absorption of waves that
occurs (when damping effects are included) at the ion cyclotron fre¬
quency is used as a means of heating a plasma in some fusion devices.
Problems
3.13 Prove that there is a small shift in the reflection points co0i and w02
when we take account of ion motions and that, from Eq. (3.6.14),
the reflection points occur at
Note that this ensures that w0i exceeds coci and a>o2 exceeds |a>ce|.
3.14 For propagation at an arbitrary angle relative to B0, the resonances
in a fully ionized cold plasma are found to occur at
w2(w2 — |cocc|ojcj) -1
(w2 — wce2)(o>2 — Wd2) _
Use this formula to make a rough plot of the three resonant fre¬
quencies as a function of y with w« > |wc|. Your result will be
138 CONTINUUM EQUATIONS FOR A PLASMA
much like Fig. 2.10, but with two exceptions: (1) the curve that
starts at |u>c| falls only to wc(wi2 + ojCi2)1/2/(we2 + uce2)1/2 as 7 >90 ,
and (2) there is the third resonance, which decreases from coci
when 7 = 0 to zero when 7 = 90°.
3.15Starting with Eqs. (3.5.5) and (3.5.7), show that when collisions
are neglected there are only two resonances in a warm fully ionized
plasma and that these are at w = |wc«| cos 7 and w = uci cos 7.
Continuity equation
Thus far in this chapter we have been using separate equations for
each type of particle in the plasma. We have always used the set of
continuum equations for the electrons, and in some sections (when the
motion of the ions was not neglected) we also used the continuum
equations for the ions. Moreover, it is clear that we could have
included the possible motion of the neutrals by adding a set of con¬
tinuum equations for the neutral molecules. An alternative approach,
which we now consider, is to replace the set of individual-species equa¬
tions by a set of equations for the gas as a whole. These new equations
are known as the magnetohydrodynamic (mhd) equations, and in their
most general form they are entirely equivalent to the set of individual
fluid equations we have been using. However, under some conditions
(generally for steady-state or slowly varying problems) the mhd
equations are particularly convenient to use and lead to results which
are not easily obtained from the multifluid equations.
Since they are entirely equivalent, we shall derive the set of mhd
equations from the multifluid equations we have already used in this
chapter. In linearized formf these multifluid equations (3.2.1) to
(3.2.4) are
where p'a, u«, pa, and E are all assumed to be infinitesimal disturbances
and pao is constant. To obtain the mass-conservation (or continuity)
f To make this derivation from the nonlinear equations one must modify
slightly the pressures and temperatures in the multifluid equations to get a simple
set of mhd equations. This point is discussed more fully in Sec. 3.10 and in Chap. 5.
3.7 LINEARIZED MAGNETO HYDRO DYNAMIC EQUATIONS 139
equation for the gas as a whole, we sum Eq. (3.7.1) over a to have
X Pa + X PaoV • ua = 0 (3.7.3)
Now, we let p and u be the density and velocity for the gas as a whole
such that
P — Po + p (3.7.4)
and
u
2 PaOUa
PO
(3.7.5)
[This last equation is, of course, the linearized form of Eq. (3.2.13).]
From Eqs. (3.7.4) and (3.7.5) it is evident that
P d- p0V • u = 0 (3.7.7)
Momentum equation
In similar fashion to obtain a momentum equation for the gas as a whole
we multiply Eq. (3.7.2) by pa0 and sum over a to obtain
Let us now consider each of these terms in their turn. From Eq.
(3.7.5) the term on the left equals p0u. Next, in the Lorentz force
term on the right
and
(3.7.10)
For almost all problems, the undisturbed or equilibrium gas has no net
charge density; hence the sum in Eq. (3.7.9) vanishes. In the pres¬
sure-gradient term we let p be the total pressure (p = 2pa) so that
^ Vpa becomes just Vp. Finally, in the last term we recall from
Eq. (3.2.14) that this double sum vanishes. Thus, when we simplify
each term in Eq. (3.7.8), we reduce the momentum equation for the
gas as a whole to the simple form
pou = J x B„ - Vp (3.7.H)
and
mj (3.7.14)
U; = u +
en0(mi + me)
mi J (3.7.15)
u
en0(mi + me)
3.7 LINEARIZED MAGNETOHYDRODYNAMIC EQUATIONS 141
mj m,;J
u — E + u X B0
en0(m,i + me) m, en0(m,i + me)
- ^ — J (3.7.16)
Pe 0
nQe2 n0e2
co w0 (3.7.18)
mvei mat0
whenever J on the left and the entire bracketed term on the right can
be neglected. Hence, Eq. (3.7.17) is also often referred to as a gen¬
eralized Ohm’s law. The bracketed term in Eq. (3.7.17) can be put
into an alternative form, if we eliminate J x B0, by using Eq. (3.7.11).
This leaves, for the bracketed term,
V x B = mo(€oE + J) (3.7.21)
so that
(3.7.23)
J err
with cr and r some characteristic conductivity and time for the plasma.
Now €0 is a little under KR11 farad/m, while for most mhd problems
a is greater than 1. Hence, unless one is dealing with very short
characteristic times, Eq. (3.7.23) shows that one can neglect the dis¬
placement current e0E in Eq. (3.7.21).
The approximations made with regard to the generalized Ohm’s
law cannot be justified in such a straightforward fashion. Rather,
it is customary simply to assume that all time variations are exceed¬
ingly small and that one has a nearly cold plasma. 1 hen, after
neglecting time-varying and pressure-gradient terms in Eqs. (3.7.17)
and (3.7.20), the generalized Ohm’s law reduces to the simple formf
shown in Eq. (3.7.19). With these approximations one obtains the
t Various ones of the other terms in Eq. (3.7.17) are occasionally used in the
mhd literature. For example, the term involving J X B0) when retained, yields
the equation (assuming p0 = m<n0 and me <<C rrii)
where the last term on the right leads to a Hall effect in mhd flow problems.
3.7 LINEARIZED MAGNETOHYDRODYNAMIC EQUATIONS 143
P + PoV • u = 0 (3.7.24)
pou = J X B - Vp (3.7.25)
J = <r0(E + u x B) (3.7.26)
V x E = —B V x B = MoJ (3.7.27)
This set still contains one more variable than there are equations,
and for many problems one must add an additional relationship
between p and p. Since we expect the plasma to behave as a simple
gas for slowly varying problems, it is generally adequate to set
Vp = U2 Vp (3.7.28)
with U the isothermal or adiabatic sound speed for the fluid as a whole.
This equation can be simply derived from the individual-species energy
equations (Vpa — Ua2 Vpa) if we assume that (as for a perfect mixture
of gases) the percentage change in each species is nearly equal during a
compression or rarefaction. Then, for example, using the adiabatic
law for each species, we have
Vp = J Vpa = y 7
*-< PaO
Vpa (3.7.29)
a a
Vp« Vp
- = - all a (3.7.30)
PaO p0
Therefore
Vp = J ypao ^ = t *1 Vp (3.7.31)
L-' P0 P0
, N (2wcT)1'2 (3.7.32)
^= 7b«
If T is given in electron volts and B0 in gauss, we hnd for electrons
(a) = 3.4 X 10->T'i*/Bo m, while for protons (ap) = IAT1I2/B0 m.
Plots of these radii are shown in Fig. 3.9. For weak magnetic helds
Fig. 3.9 Plot of cyclotron radius vs. temperature for several values of
magnetic-field strength. The solid lines are for electrons, the dotted for
protons.
3.8 MAGNETIC PRESSURE, VISCOSITY, REYNOLDS NUMBER, AND DIFFUSIVITY 145
it can be seen that the proton cyclotron radius is fairly large, and for
heavier ions it would be still larger. Hence the effect of the walls in
inhibiting the cyclotron motion of the ions must often be taken into
account.
Magnetic pressure
(V x B) x B = (B • V)B - (3.8.1)
(B-V)B / , B2\
Pou ----V [p + ) (3.8.2)
Mo \ 2po/
The term 52/2/x0, which enters here, is called the magnetic pressure
since it plays a role entirely analogous to the hydrodynamic pressure p.
For a steady-state problem where the magnetic field lines are
straight (so that B is perpendicular to any gradient in B), Eq. (3.8.2)
states that p + B2/2uo is constant. This simple result provides a
rough basis [since a true steady state and an exact cancellation of
(B • V)B is not possible] for estimating the confining fields needed in
a fusion device. Ideally, outside the confinement region p is zero, so
that if B0 and B are the field strengths respectively outside and inside
the plasma, we have
B*_ Bl B2
(3.8.3)
V + or 0 = 1 Bo2
2mo 2po
Problem
3.16(a) Show that a plasma with a pressure of 100 atm requires a
field of at least 5 webers/m2 (5 X 104 gauss) for magnetic
confinement. (This is a pressure that is adequate for a
fusion device.)
(6) Show that if this field of 5 webers/m2 is produced by a current
flowing through a cylindrical column of plasma (as in the
pinch effect) with a radius of 10 cm, then the current needed
is 25 X 105 amp. (Of course these would be curved field
lines, so this calculation is just to show the order of magni¬
tudes involved.)
-V xE = Vx(uXB)--VxJ (3.8.4)
&0
V x (V X B) = — V2B (3.8.5)
uB vmB
V X (u X B) I'm V2B (3.8.7)
17 L2
3.8 MAGNETIC PRESSURE, VISCOSITY, REYNOLDS NUMBER, AND DIFFUSIVITY 147
uL
Rm (3.8.8)
Vm
with f the average external force per unit mass and vk the kinematic
viscosity (viscosity divided by density). When we compare this with
Eq. (3.8.6), it is clear that vm plays essentially the same role in the
decay of B that vk does in the decay of the fluid velocity u. Here the
Reynolds number is the ratio of the inertia term u • Vu to the viscosity
term vk V2u or, by dimensional analysis,
u2/L _ uL
(3.8.10)
vku/L2 vk
B vmB
B - and vm V2B (3.8.12)
TD U
148 CONTINUUM EQUATIONS FOR A PLASMA
SO that
IS (3.8.13)
TD — -
Vm
JB • dS = Jv x (u X B) • dS = f(u X B) • d\ (3.8.14)
we have
JB • dS + /B • (u x dl) = 0 (3.8.16)
Now we shall show in the next paragraph that for a surface moving
with velocity u
Hence, Eqs. (3.8.16) and (3.8.17) show that for the moving surface
JB • dS = const (3.8.18)
Thus the flux passing through the moving surface stays constant.
This is known as a frozen field condition, since the magnetic field lines
are, in essence, “tied” to the fluid and move along at the fluid veloc¬
ity u.
To complete this discussion we now want to prove the identity
given in Eq. (3.8.17). Suppose we consider a moving open surface S
3.8 MAGNETIC PRESSURE, VISCOSITY, REYNOLDS NUMBER, AND DIFFUSIVITY 149
such that, at time t, S(t) — Si, and at a time At later, S(t + At) = S2.
Then, as in Eq. (3.1.5),
d
dt
(/s.ds)
(3.8.20)
(j)B • dS = f V • B dV = 0 (3.8.21)
We now apply this to the closed surface (shown in Fig. 3.10) consisting
of Si, S2, and the “sides” of length u At, to have
(Note that the minus sign in the second term is due to the fact that
S2 points into the volume.) If we use this result in Eq. (3.8.20) and
note that
u M->>
we have
Problems
3.17(a) Show for a typical mhd generator with u = 103 m/sec,
a = 100 mho/m, and L — 0.1 m that Rm « 1 and the
diffusion time is very short. (Hence inhomogeneities in the
field are quickly smoothed out.)
(6) In a fusion device (where inhomogeneous confining fields
should be maintained for a fairly long time) show that diffu¬
sion times on the order of 103 sec are possible when Te is of
the order of 104 ev and L ^ 1 m. Note that at these tem¬
peratures, as indicated in Eq. (2.2.17), ao = 4.6(T/300)3/2,
with T in degrees Kelvin. Show that, for such a device, the
field lines are essentially frozen.
3.18 Derive the equation pou = a0(E xB) + a0(u X B) X B - Vp,
and show that for E = 0 and p constant the velocity perpendicular
to B is given by ux = ux(0)e~tlT, with r = po/oqB2 = vm/\ a2.
Calculate r for a fusion device with n = 1022 protons/m3,
B = 5 webers/m2, amd T = 104 ev.
P d- Po^ • u — 0 (3.9.1)
and
/ P0 /| \
p = — (k • u) (3.9.4)
CO
and
Va = (MOPo)-1/2Bo (3.9.8)
Hence, in the last term on the right side of Eq. (3.9.9) we have
A = u X V„ (3.9.12)
k X (k x A) = — k2A, (3.9.13)
SO that
From the last of these equations we see that there can be a wave
linearly polarized in the 2 direction (the direction perpendicular to
both k and B0) provided
or
The solutions with the plus-and-minus sign here are (for obvious
reasons) called fast and slow mhd waves.
All three of these mhd dispersion relations (3.9.18) and (3.9.21)
describe waves with constant phase velocity for all frequencies. Hence
154 CONTINUUM EQUATIONS FOR A PLASMA
Problem
3.19 Note, from Eq. (3.9.7), that we can formally include the effect
of vm by replacing Va2 by
Va2
(V'a)2 1 + i(vm/ w)/c2
Use this fact to find the damping factor for each of the mhd
waves when vm is very small but nonzero.
Fig. 3.12 Plot of phase velocity as a function of the angle y between the
direction of wave propagation and B0 for the three types of mhd waves.
3.10 THE NONLINEAR MHD EQUATIONS 155
Pa + V • (paua) = 0 (3.10.1)
Duak
Pa naq«[E + (ua X B)]*
Dt
p + V • pu = 0 (3.10.3)
where p and u are again the total density and total fluid velocity as
defined in Eq. (3.2.13). This simple result has precisely the form
one would obtain for a simple fluid with density p and fluid velocity u.
In similar fashion, if we sum the momentum equations for all the
species, we can obtain ultimately an equation of the form expected
for a simple charged fluid,
Duk dp0
PcEk + (J X B)fc (3.10.4)
p Dt dXk
(Note that this extra term involving the diffusion velocities is of second
order in wa; hence it drops out of the linearized mhd equations.)
To derive Eq. (3.10.4) we first sum the individual momentum
equations (3.10.2) over all a. Then, from the definitions of pe, J, and
dp°/dxk and the identity [shown in Eq. (3.2.14)] that
^ PaVaffUa ~ Up) - 0
(3.10.7)
a,0
we have
duak
Uak "T Uai —-
i )
To simplify this equation we must go through some algebra and use the
identity
(3.10.9)
£ PaVfa = X P“(U“ ~ U) = PU _ PU - 0
(3.10.10)
(3.10.11)
dt
The second term here is identically zero [from Eq. (3.10.3)], and the
third term exactly cancels the last term on the right side of Eq. (3.10.8),
so that we obtain the simple nonlinear momentum equation
(3.10.15)
(3.10.16)
ep
ynlpi
Dui
In mePe T)f j
D uA
E + u X B — — (m*
ep
Vpi me V pe)
The terms on the left are the usual ones retained in magnetohydro¬
dynamics, while all the terms on the right are generally discarded.
Clearly, the omission of all these terms is not always justified, hence,
one should check the magnitude of these terms to see whether they
are truly negligible in a given mhd problem.
chapter four
THE BOLTZMANN-
VLASOV EQUATION
FOR A PLASMA
chapter four
dNa (4.1.1)
/a(x,v,i)
d3x d3v
(4.1.4)
(4.1.5)
Fig. 4.1 The volume element d3x and the velocity-space element dh.
162 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
f Note that in this equation and elsewhere we omit the subscript next to an
average when it is redundant, that is, (ca) = (ca)«.
f With this definition Ta equals the absolute temperature in thermodynamics.
§ Note that, to make this calculation, we assume that dvk At is negligibly small
and that collisions during the time interval At are also of negligible importance.
4.1 DISTRIBUTION FUNCTION AND THE BOLTZMANN EQUATION 163
Hence the flux T(Q), which is the rate of transport of Q per unit time
and area, is just
= Pa(cafak) (4.1.12)
and
r i0 /c 2C \ (H. 1.1.5;
yak "2Pa\^a ca«7
Problem
4.1 For a gas mixture one can define an alternative random velocity,
c° = v — u, where u = 2paii0/2pa is the velocity for the gas as a
whole. Similarly one can define an alternative temperature
Ta°, pressure pa°, and so forth simply by replacing ca in Eqs.
(4.1.9) and (4.1.12) to (4.1.14) by c°. Prove that
Vaik — P ~P PaLCaiU'ak
dP°aik dpa
dXi dXk
+ dx
Pa^ax^ak
df* dfa
dt
+ v■
dxi
+ di (4.1.17)
Problem
4.2 In deriving the Boltzmann equation we have assumed that the
elements d3x' d3v' and d3x d3v are equal. Actually, from the
theory of Jacobians, we have
General equation
In the preceding section we saw that macroscopic variables such as ft,
u, T, and p which enter the continuum equations can all be calculated
from the distribution function fa. The distribution function, in turn,
satisfies the Boltzmann equation (4.1.17) derived above. Hence, as
one might expect, it is possible to derive all the continuum equations
(like the mass, momentum, and energy equations used in Chap. 3 or
the viscous-stress and heat-flow equations we shall use in Sec. 5.7
and 5.8) from the single Boltzmann equation. In this section we shall
first derive the general transport (or continuum) equation for an
arbitrary function Q, and thereafter we shall use three specific Q’s
which lead to the mass, momentum, and energy equations.
To start, we have the Boltzmann equation
dzv ~ f fa ^ d*V
I = £/ <*• dt
/ Vi^ Qd3v
dx;
dQ\ (4.2.3)
na(Qvi)a na (Vi
dXi/a
dQ
d3v (4.2.4)
f Note that this restriction does not rule out the force due to a magnetic field,
a = (q/m)(v X B), since one can easily verify that a; is still independent of vx.
[For example, ax = {q/mfivyB, — vzBv), which is independent of i^.l
4.2 TRANSPORT EQUATIONS 167
Now consider the integration over dvi in the first integral on the right.
We have (omitting the subscript i throughout) for one of the integrations
(4.2.5)
with
•o--/0®.* <4'2'8)
Mass-conservation equation
Equation (4.2.7) is a general result which holds for any arbitrary
function Q(x,v,t). To derive the mass-, momentum-, and energy-
conservation equations, we let Q — ma, macak, and \maca2, respectively.
Thus for the mass equation, with Q = ma, we have
by collisions, and
(4.2.10)
[This result also arises very simply when we use the Boltzmann expres¬
sion for (dfa/dt)c derived in Chap. 5.] Hence, for this case if we sub¬
stitute the expressions in Eq. (4.2.9) into the general transport equa¬
tion (4.2.7), we find
d . d (4.2.11)
= 0
ft p“ + a5
pa + V ■ (PaO = 0 (4.2.12)
Momentum equation
To derive the momentum equation, we let
where dik is the Kronecker delta, which equals one for i = k and zero
otherwise. Using these identities, we have for each of the terms in the
transport equation (4.2.7)
j-na{Q)a = 0
Ha(Q)0: — pa'dak
Ha \ T \ = Pa (di$ik)a Pa(P'fc)a
\ OVi/a
4.2 TRANSPORT EQUATIONS 169
(4.2.16)
where
(4.2.17)
The first of these follows at once, since the average external force
per unit mass due to electromagnetic fields is
— (E + v x B)a = — (E + u„ x B) (4.2.19)
ma ma
Condition 2, however, follows for a gas only when viscous effects are
negligible. This will be clarified in Chap. 5, when we show that
Energy equation
To derive the energy equation we again work directly with the general
transport equation. We let Q = -|maca2 and note that
na(Q)a = #P« Q — a • ua
(4.2.21)
dQ dUaj dQ_
= 'ffl'dCaj = macai
dXi dXi dVi
Using these identities plus those in Eqs. (4.2.14), we have for each of
the terms in Eq. (4.2.7)
d , d
jtMQ)a - dtsVa
PaX^-a * (^a) ^
d d d
— na(QVi)a = — |pa(ca2(ca, + Uai)) = (?«* + %PaUai)
dXi dXi OXi (4.2.22)
The last term here goes to zero, since for a velocity-independent force,
(a,iCai) = a,i(cal) = 0, while for the force due to a B field (the only
common velocity-dependent force)
where 8Qa0 is the collision term, fQa(dfa/dt)c d3v. We can rewrite this
in a form more reminiscent of the adiabatic energy equation used
earlier by noting that the last two terms on the left side of this equa-
4.2 TRANSPORT EQUATIONS 171
S 11 a j duaj
V 2Pa^-a 1
~“ Paij 2P* Ua ~\~ 2'Ua * Vpa + Paij (4.2.25)
dXi dXi
V • ua = - - ^ (4-2.27)
Pa Dt
so that, from the terms on the left side of Eq. (4.2.26), we have
DTa dUaj
(4.2.30)
^nan “h N • CJa ""I- Paij 8QaO
Dt dXi
which reduces to a diffusion equation for Ta when ua and 8Qa0 are zero
and when qa = — X VTa, with X the thermal conductivity.
172 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
Problem
4.3 Derive Eq. (4.2.30) for the temperature. When the equation
reduces to a diffusion equation, show that if X = 5i<pa/2mava,
then the characteristic diffusion time is given by
3 L2
T 5vth2
SUMMARY
In this section we have seen that just as most macroscopic variables
can be computed from fa, so too the equations for the time rate of
change of these macroscopic quantities can be found from the Boltz¬
mann equation. An important point (which is evident in the three
transport equations we derive here) is that the set of transport equations
derived from the Boltzmann equation always includes more unknown
functions than equations. Thus the equation for the time rate of
change of pa includes ua; the equation for the time rate of change of
ua includes pa; and so forth. Any finite set of transport equations is
therefore always insufficient to provide a closed set of equations for a
problem, and it is necessary to introduce some type of approximation
(or truncation) scheme to eliminate some of the functions or to express
some of the functions in terms of others. A systematic approach for
working with the transport equations is discussed in Chap. 5.
In the remainder of this chapter we shall describe some problems
where we work directly with the Boltzmann equation itself and solve
for the distribution function. In principle this is always the most
satisfying technique, since one can calculate any macroscopic variable
once fa is known. However, as might be expected, the Boltzmann
equation is really too difficult to solve in general, and before we can
obtain a solution to any problems of interest, many approximations
must be made. It is for this reason that the transport equations
continue to play an important role even in kinetic theory.
The BV equation
We have seen in previous chapters that although a plasma is a mixture
of electrons, ions, and neutral molecules, one can explain many phe¬
nomena fairly well by considering just the motion of the electrons.
4.3 THE BV EQUATION AND THE RELAXATION MODEL 173
a = - - (E + v X B) (4.3.2)
m
with E and B the slowly varying fields which enter Maxwell’s equations.
The first decision to be made at this point is how to treat the
collision term in Eq. (4.3.1). The simplest assumption is merely to
neglect collisions and set
+ + = ° {4-3'3)
©.--'tf"" . . (4'3'4)
or
(4.3.6)
/ + vf = vfo
/ = ce~vt + fo (4-3‘7)
which shows that the difference between / and /0 decays to zero expo¬
nentially at a rate proportional to v.
To summarize, the Krook (relaxation) model provides a very
simple collision term which tends to drive / toward some equilibrium
distribution /<, at a rate governed by the relaxation time v~l. This
seems, intuitively, to agree with the role one might expect collisions
to play, and, moreover, as we shall see in Chap. 5, there are many
examples one can cite where the model leads to almost the identical
results obtained by using the full Boltzmann collision term.
Although we shall use the Krook collision model primarily for problems
where we solve directly for the distribution function /, it is also of
interest to note that the model greatly simplifies the evaluation of all
the collision integrals that enter the moment equations. From Eq.
(4.2.8) recall that
•«-/«©* (4'3'9)
Hence, using the Krook model for (df/dt)c and the definition of an
average in Eq. (4.1.4), we have (for v independent of velocity)
where
9 = ffQ dh = n(Q)
(4.3.11)
go = SfoQ d3v = n(Q)o
For a simple gas, i.e., a gas having only one species of particles, the
mass, momentum, and energy densities cannot be changed by collisions
(because of the conservation of these quantities in elastic collisions);
hence/0 must satisfy the three conditions
m Jfo dh = p
mffov dh = pu (4.3.13)
±mjf0v2 dh = ~2-p(t’2) = + *-pu2
where p (or n), u, and T are the true density, mean velocity, and
temperature in the gas.
A common choice for/0 is the local maxwellian velocity distributionf
ne' ~c2/a2 2 kT
fo = c = V — u, (4.3.14)
7r3/2a3 m
One can easily verify that this function satisfies all three conditions in
Eq. (4.3.13), so that 8Q is zero in the mass, momentum, and energy
equations. [The integrals involving f0 in Eq. (4.3.13) are easily
handled if one changes the variable of integration from v to c — u
with dh = d3c.\
u(t) = u(0)e vt
(4.3.17)
±m(v2)t — f«T0 = (|m(v2) 0 — ^Kl\)e~vt
with (v2)t the average value of v2 at time t. Thus, in accord with Eq.
(4.3.16), the Krook model causes both the drift velocity and the mean
energy to approach their equilibrium values in a time of order v~l
(where v for this case would be the electron-neutral collision frequency).
In fact, however, an analysis of the collision process using a more
detailed collision model reveals that while the equilibration time for
the mean velocity is approximately v~l, that of the mean energy is
much longer, approximately M(2mv)~1, with M the neutral-molecule
mass. This is easily understood by remembering that an elastic
collision between a fast electron and a slow neutral molecule is much
4.4 APPLICATION TO LONGITUDINAL ELECTRON WAVES 177
Av
v
« 1 (4.3.18)
Hence, after a time long enough for an average electron to have one
collision with a neutral molecule, we can expect the average velocity
of the electrons to change significantly but not their average speed
(or kinetic energy). For this reason, the Krook model in this simple
form is really applicable only to cases where self-collisions are of
predominant importance.
so that the wave travels in the x direction. Substituting this into the
BV equation (4.3.3) and linearizing, we have
so that
__ ieE(dfo/dvx) (4.4.3)
1 mk[(w/k) — Cx]
178 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
V • E = - = e<yUi ~ ^ (4.4.4)
Co fo
High-phase-velocity limit
b
dV = UV
a }>dU (4.4.8)
where
U -
(4.4.9)
dV = ^ dvx V = f0
dvx
4.4 APPLICATION TO LONGITUDINAL ELECTRON WAVES 179
The first term on the right is clearly zero since f0 vanishes when
vx = +oo. Hence the full dispersion equation (4.4.7) reduces to
(4.4.11)
(4.4.12)
co2 = co 2 (4.4.15)
[It will be recalled that this is also the cold-plasma result shown in
Eq. (2.4.27).] To consider small corrections to this result consider the
next two terms in the dispersion equation. First, if the plasma is
180 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
(4.4.16)
(vx2)o — (Cz2)o — -g-(C“)o m
= . • •) (4.4.17)
y m ad /
To compare this with previous results, note that the underlined term
here is very small (since w/k has been assumed to approach infinity),
hence, we can replace w in just this small term by we (the value for w
when the term is zero) and obtain
2 + m
7 = (4.4.19)
m
Problem
4.4 Use the BV equation (for each species) and Poisson’s equation
to obtain the following dispersion equation for small-amplitude
longitudinal waves in a fully ionized gas with equal ambient
4.5 LANDAU DAMPING 181
i = i r/ \ . / \
k2 _\(vx — co/k)2/eo \(yx — CO//C)V,0_
Now we shall show how this equation can be evaluated for the impor¬
tant case when /0 is the maxwellian velocity distribution for a station¬
ary plasma (u =■ 0), so that, as in Eq. (4.3.14),
n0e~v2lai 2 kT
/o =
(4.5.2)
7r3/2a3 m
/ = /o + and E =
reduces to
e Fdf0 (4.5.3)
— iu/i + ikvxfi -r— -vfi
m dvx
dfo (4.5.5)
-2^2/o
dvx
— A’d2/
(4.5.7)
_fc2"
fcD2 (4.5.8)
kT/rn
CO + iv
(4.5.9)
I = C =
ka
hence we have
(4.5.11)
4.5 LANDAU DAMPING 183
Of the three terms on the right, the first is a well-known integral which
equals unity; the second integrand is odd, and its integral therefore
vanishes; and the third integral is one which we must still evaluate.
Hence, to this point, we have
1 = 1+ C2G( 1) (4.5.12)
with
sq*
[" q2e
7T l/2 dq
/-« q2 -
Tr-i/2
( C2 \ ,
[ * ^ 1
S-112 .
- C2G(s) (4.5.14)
G(0> ^ (4.5.17)
• l/2
s^C = 2 ds = (4.5.20)
so that
1/2
t7T
2 (4.5.21)
Gd) - -§ /%>'-<’* + c
D-C
If we substitute this result into Eq. (4.5.12) for I and then apply that
result to the dispersion relation given in Eq. (4.5.7), we find
w + iv
(l - 2C foC ez2~C2 dz + i7r1'aCe-c‘) C =
ka
(4.5.22)
The integral which remains here is known as the dispersion f unction and
has been tabulated, f while the last term, which is imaginary, is known
as the Landau damping term. To calculate k as a function of w from
this dispersion relation ohe may use the following procedure:
1. Choose an arbitrary C.
2. Find the (tabulated) value for the dispersion function.
8. The dispersion relation then gives k2.
4- Find co from the definition, C = (u> + iv)/ka. (Actually, in most
calculations v is neglected, so that C = co/ka.)
11 = 2 C fj ez2-C2 dz
where Oe~°2 denotes terms of order exp( — C2). The derivation of this
expansion will be given shortly, but assuming that it is valid, we can
use it in Eq. (4.5.22) to find
Now C2 = w2/k2a2 (in the limit when we set v — 0), and kD2 = 2o>e2/a2.
Hence this dispersion relation easily reduces to
Landau damping
Again, since C has been assumed to be very large, we have, to lowest
order, w2 = a>e2, and
The first two terms here are the Bohm-Gross result [Eq. (4.4.18)].
However, the imaginary term in the dispersion relation is new. Note
that for a standing-wave problem (where k is real) this result shows
that w2 has a negative imaginary part, so that, from the arguments in
Sec. 2.4, there is temporal decay. This is also evident here since, from
Eq. (4.5.27), to lowest nonvanishing order in real and imaginary parts,!
iirll2coe3 (4.5.28)
e we — ia
k3a3
The negative imaginary part in o> leads to temporal decay, since the
waves are proportional to
(4.5.32)
(4.5.33)
with
, n\ n n(n — 1)
2 + +
(
-C
e x dx £f2n e 1 C2
~C^
(4.5.35)
can use the first few terms in the expansion to obtain a good estimate
of I\ provided C is large.
Problems
4.5 Prove that, for small values of C,
so that, as C —* 0, k2 = — ko2 or
k7
- = —03,
m
Analysis with B0 = 0
To find the general dispersion equation for small-amplitude waves in
the electron gas we set
dfo = dv_dfo =
(4.6.2)
dvi dvi dv v 0
|Tpikx—uat (4.6.3)
Note that the magnetic-field term does not enter, since, from Eq.
(4.6.2),
(v x B)ip- = (v x B) • = 0 (4.6.6)
dVi V
kax X E = wb (4.6.8)
f For some problems this assumption cannot be used, and the resulting analysis
is somewhat more complicated.
t Note that, again, if we use the Krook collision term, the only change is from
to to a; + iv in this equation. For this reason, where needed, we shall evaluate all
poles as if we had used the Krook model.
190 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
J = —ejfi\d3v (4.6.10)
E =_
n0co(l
^E-—~ n2) Jf a/°
_
dVi co
Vi dh -
— kvx
(4.6.13)
Two of the terms vanish in each sum on the right side of Eqs. (4.6.12)
and (4.6.13) since they lead to integrals of the form
r vx(dfo/dvx)
J kvx — co
d,v
This equation has been analyzed thoroughly in the last two sections
and need not be discussed further here. However, from Eq. (4.6.16)
we see (after an integration by parts) that for transverse waves (Et 9^ 0)
we have the new dispersion relation
coe2 ■ r fo d3v
(4.6.19)
1
n0co(l — n2) J kvx — co
n0e-”2/a2 (4.6.20)
7r3/2a3 m
l = ~W*2C_ (4.6.21)
7t1/2co2(1 — n2) Q~C
CO e2C2 (4.6.22)
1 = G(l)
co2(l - n2)
where G(s) is the integral defined in Eq. (4.5.13). In the limit when
C is very large we have seen [Eqs. (4.5.21) and (4.5.23)] that
1/2
ITT
2
i( 1+J_ + ± + + c -C
(4.6.23)
(7(1) c! V ^ 2C* 4C* T )
192 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
Bn + beikx~iat (4.6.25)
(■e/m)2g(v)Ei
/i(1) kvx){\ X B0)i + kvi(\ X B0)J (4.6.31)
B„(« - kvxy u
Hence, to obtain our dispersion relations (to first order in B0) we need
only substitute our approximation /i = /ico) + B0f i(1) into this equa¬
tion and carry out the integrations over v. If we let
so that after carrying out the integrations over vy and vz, we can reduce
Eq. (4.6.32) to the form
A•E = 0 (4.6.34)
where
2C r q2e~gl r _ C r er* dg
Ti
J
7T1/2 ' g-cd? 12 ~ ir112 J q-C
T3 = C2 AA
dC c
_ ifiZ At A (4.6.37)
t4
2 dC2 C
t5 — C2 A. ^1
= dC~C
_ 2C2 r” qe~q'1
dg
7J — » q — C
= 2C2[1 + C2G(1)]
r2 = <7*G(i)
4.6 LANDAU DAMPING OF TRANSVERSE WAVES 195
so that A reduces to
1 - X 0 — iXY sin y
A = 0 1 - n2 - X iXY cos 7 (4.6.39)
iXY sin y —iXY cos y 1 - n2 - X
with
sin2 y
B = X2F2 C = X Y cos2 2 2 7 (4.6.41)
1 - X
Hence we have
n2 = 1 — X - ± (}B + C)1/2
2 (4.6.42)
n 2 = 1 - X ± XF cos 7 (4.6.43)
n = i-—-^ 1 — X ± XF cos
2 7 (4.6.44)
1 ± F cos 7
However, this is not surprising in view of our having kept only terms
of order B0 (or Y) in the expression for /j.
Hence
dfi^dVxdfi.dVydfi.dtudfi
d<f> d(f> dvx d<t> dvv d4> dvz
(4.6.46)
y
Fig. 4-4 Coordinate sys¬
tem with az in the direc¬
tion of B0 and ax perpen¬
dicular to k and B0.
4.6 LANDAU DAMPING OF TRANSVERSE WAVES 197
Through the use of this identity, the BV equation now reduces to the
form [see Eq. (4.6.27)]
/ | \ r ^ X? 0 I * d/1
(co - k • vi = — Ei --h iuc — (4.6.47)
m di’i d4>
f To take the derivative of fi recall that the formula for the derivative of an
integral is given by
dG ,. G(x + *) - G(x)
—— = lim -
dx e_^o e
J See for example, I. B. Bernstein, Phys. Rev., 109:10 (1958) or E. G. Harris,
J. Nucl. Energy, pt. C, 2:138 (1961).
198 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
Problem
4.7 Perform the integration indicated in Eqs. (4.6.48) and (4.6.49)
for waves traveling in the direction of Bo (so that ky = 0) to find
ie 1
h = A2 - 1
mwc
Choice of /0
written as
where the 8 functions shown here are discussed in Sec. A.l of the
Appendix.
fo dh (4.7.2)
0 = w + iv
n0 J (.kvx — ft)2
1 1
_(kV - ft)2
+ (.kV + ft)2
(4.7.3)
or
We can solve this equation either for standing waves (where k is real)
or traveling waves (where a> is real). Since it is somewhat easier with
the Krook model to treat traveling waves, we rewrite Eq. (4.7.4) in
the form
j The use of the simple relaxation model is not really justified here, since fo
is not a maxwellian distribution and collisions do not tend to damp out deviations
between / and/o. However, as in the last sections, we shall always make our cal¬
culations in the limit as v —* 0.
200 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
In this limit the real and imaginary parts of k are both positive for
each root; hence both waves are slightly attenuated. The phase
velocity for each wave approaches the beam speed V as w approaches
infinity, but for finite frequencies the phase velocity is less than V for
the root with the plus sign and greater than I for the root with the
minus sign, as shown in Fig. 4.5.
The growing wave appears when we examine the low-frequency
limit, where |122| « we2/8, so that Eq. (4.7.6) reduces to
2122
k2V2 ^ iue2 1 + — ± (4.7.9)
U>e
Since the imaginary part of k2 is negative for this root, from the dis¬
cussion in Sec. 2.5 it follows that this is a growing wave. It is worth
emphasizing the importance of using the Krook relaxation term.
Had we used just the collisionless Boltzmann equation, the dispersion
Fig. 1^.5 Plot of phase velocity and growth rate vs. frequency for waves
traveling in two contrastreaming plasmas.
4.7 THE TWO-STREAM INSTABILITY 201
relation here would have been k2V2 = — co2, and we should not have
been able to demonstrate clearly that the wave is growing rather
than evanescent, f
In Figure 4.5 we show a plot of phase velocity and growth factor
(the imaginary part of k) as a function of frequency for these two waves
in the contrastreaming plasma beams in the limit when v = 0. As
shown in the figure, the root with the minus sign grows for all fre¬
quencies below the plasma frequency. This can also be seen from
Eq. (4.7.6), where, for v = 0, the root with the minus sign is of the
form
k2V2 = A - B (4.7.12)
where
and
B2 = 2coW+iwc4 (4.7.14)
so that, for co2 <C co,.2, B exceeds A and k2 is negative, dhe maximum
growth rate occurs when co2 = f coe2, as can be verified by examining
the derivative of k2 with respect to co. 1 his amplification effect has
been verified experimentally and serves as the basis for the two-
stream amplifier.
Problems
4.8 Derive Eq. (4.7.3) from Poisson’s equation and .the macroscopic
cold-plasma mass and momentum equations for two beams
of electrons where 711,2 = + n\:2elkx^l0,t, v\x = I + vxelkx lat,
V2x = —V A- v'2eikx~iat, Ex = E'eikx~io,t, and V » v'. Note that
one should neglect collisions; hence the result will have co rather
than H.
4.9 Verify that the maximum in the growth rate occurs when
w — (3)i/2w<, What is the number density, mean velocity, and
f We should note that this conclusion, though it follows directly from the use
of the Krook model, does not agree with an analysis (using the BV equation) by
P. A. Sturrock [Phys. Rev., 112:1488 (1958)], who concludes that for contrastream¬
ing plasmas there is only temporal growth and not spatial growth. However, this
difference is probably only a question of the choice of a moving or stationary coor¬
dinate system, as discussed by R. J. Briggs, “Electron Stream Interaction with
Plasmas,” The M.I.T. Press, Cambridge, Mass., 1964.
202 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
temperature for each beam and for the mixture of the two beams:
[Find these by working directly with fo as given in Eq. (4.7.1).]
Wb2 , C0p2
(4.7.16)
(kV - ft)2 + TF
where co6 = (nbe*/me0)1/2 and cop = {npe2/mt0)I/2 are the plasma fre¬
quencies for the beam and for the stationary plasma, respectively.
From Eq. (4.7.16) we have
- e>! - ^ (47-I7)
Hence, if we take the square root of each side and go to the limit as
v —> 0, we have
COb
kV = co 1 + (4.7.18)
(co2 - CO,2)1'2
When co exceeds cop, k is real, so that the waves are neither damped nor
growing. However, for co below cop, we have
IWb
kV = CO 1 + 1/2
(4.7.19)
(cop2 — CO2)
so that the phase velocity co/Re k is V for both roots, but one root is
damped, and the other grows at the rate
COCOb
Im k = (4.7.20)
F(cop2 - CO2)1'2
4.8 THE DEBYE POTENTIAL PROBLEM 203
CO
Fig. 4.6 Plot of phase velocity and growth rate vs. frequency for wave travel¬
ing in a beam-plasma amplifier.
204 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
E = -V4> (4.8.1)
and
e d4> dfi
1) ■ dfi- (4.8.3)
1 dxj mi dxj dVj
To these two equations which involve fe, fi, and 4> we again add
the Maxwell equation involving V • E (Poisson’s equation) to have
Pc
V • E = (4.8.4)
«0
V24> = —
£0 J
f
(fe — fi) dzv — — 8(x)8(y)8(z)
£0
(4.8.5)
\mav2 + qa$\
fa = A a exp (4.8.6)
«T )
where a takes on the values e for electrons and i for ions and where Aa
is a constant. We can easily verify that this satisfies Eqs. (4.8.2)
v dfa _ dfa
- 1) = 0 (4.8.7)
3 dxj ma Ox, dvj K 1 OXj
N = jj fad3vd3x
N • . N
A, — Ai =
7T 3l2ae3f ey d3x IT 3/2a,i3fe y d3x
(4.8.9)
e<t>
T rn
Kl
where
ey e4>
7 = (4.8.12)
W) = fey d3x fe~y d3x kT
206 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
F(4>) =
1+7
7(1 + 5)
1 7 =*
7(1 - 5) ~ 7
% (7 - 5) (4.8.13)
Even with this simplification, the equation for <t> is still too complicated
to treat analytically unless we also have
2co„2 = 2(N/V)e2
(4.8.18)
kT/wi K.TtQ
To complete our problem we must now solve Eq. (4.8.17) for the
potential 4>. Since the source of the potential (the point charge Q) is
spherically symmetric and located at the origin, <t> should have no 6
or </> dependence. Moreover, in spherical coordinates, when 4> depends
4.8 THE DEBYE POTENTIAL PROBLEM 207
l R$ - kD'^ = 0 (4.8.20)
R dR2
or
— R® ~ kD-2m) = 0j (4.8.21)
aKz
To evaluate the first term on the left we first apply Gauss’ theorem
to obtain
(4.8.26)
dS = t2 sin d dd d<\> ar
208 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
while
f
J R <(
V24> dV = -
J0
r B [exp ( — kD>e)](kD>e + 1) sin 6 dd d<t>
JO
-> -47rB (4.8.28)
while, for the term on the right, the integration over the 5 functions
yields unity, so that the entire equation reduces to
Q exp( — kD'R)
(4.8.31)
47veoR
Note that if the point charge Q were placed in a vacuum, coe (and
therefore kD') would be zero, so that the potential would reduce to
the familiar coulomb potential, 4>c = Q/^-k^R. Therefore, we can
4.8 THE DEBYE POTENTIAL PROBLEM 209
which shows that <t> is much less than the ordinary coulomb potential once
R exceeds the Debye length. Hence, in a crude way, one can say that
a charged particle in a plasma interacts (in steady state, at least) only
with particles less than one Debye length away. These particles are
said to lie in the Debye sphere surrounding the charged particle.
To complete our discussion of this important problem, we now
can reexamine the two approximations used to obtain the Debye
potential: (1) that |y| = \e$/kT\ <3C 1 and (2) that
To check the first approximation, note that with <f> given by Eq.
(4.8.31) and Q = e,
Problems
4.10 Verify Eq. (4.8.34) by evaluating 6 = V-'jy d3x over a sphere
of radius E0 and letting <t> be the Debye potential [Eq. (4.8.31)].
4.11 Using the collisionless hydrodynamic momentum equations for
the electrons and ions
Pa
I)U a __
^Pa
but with
(4.9.2)
we find
na rx r°° r°
ra = vxe~v*laa2 dvx dvy dvz
7r3/2a«3 J-« J -» yo
na r« (4.9.3)
vxe~Vi,,aa2 =
7r1/2aa jo
From this calculation it is evident that if initially the plasma near the
wall has equal number densities for the electrons and ions, then I\>
far exceeds I\ as a result of the large difference between (Te/me)1/2 and
(Ti/m,i)1/2 in most plasmas. For this reason, a wall in contact with a
plasma rapidly accumulates a negative charge.
The potential associated with this negative charge provides an
attractive force for the ions and a repulsive force for the electrons in
the plasma, so that I\ increases and 1% decreases, until eventually an
equilibrium is established where I\ and re are equal and the negative
potential at the wall has a fixed value.
„ V± . A df? =o (4.9.4)
1 dx me dx dvx
4.9 PLASMA SHEATHS 213
and
V*J-
ox - —TTY1
rrii ox ovx = 0 (4-9.5)
In addition, since the charge density in the plasma is just that due to
the electrons and ions, Poisson’s equation has the form
V • E = - (4.9.6)
or
d24>
— (ne — rii) ~ [ (fe ~ fi) d3V (4.9.7)
dx2 to to J
n0 2kT
aa2 (4.9.9)
U =
Tr3l2aa3 ,T)
=
ma
214 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
where the constant in front has been set equal to no so that the electron
and ion number densities both equal n0 in the interior of the plasma,
where <t> = 0, and where we have also assumed, for simplicity, that
Te = Ti. We should note at once that several objections to using
Eq. (4.9.9) can be raised. First, for this distribution function the
probability of particles’ moving toward or away from the absorbing
wall is equal; that is, fa(vx) = fa{-vx). However, for the ions in
particular it is evident that near the wall almost all particles must
be moving to the right (since no charged particles are emitted by the
wall). Similarly, the mean velocity, as calculated from Eq. (4.9.9),
is zero. Yet, again we expect that there should be a net drift of both
electrons and ions toward the wall in order to replenish the charged
particles that are lost there.
Because of these inadequacies we shall not attempt to use Eq.
(4.9.9) for fe and /, in Eq. (4.9.7) and thereby obtain a differential
equation for 4> alone.f (This is left to Prob. 4.14.) However, we
shall use this distribution to obtain an approximate value for the
potential on the wall. First, from Eqs. (4.9.1) and (4.9.3) we have
[using Eq. (4.9.9) for/e and fi]
(4.9.11)
or
(4.9.12)
f The error involved in such an analysis is often not so bad as one might fear,
since Eq. (4.9.7) involves the integral of fe and/,- over all v, which is not too sensitive
to the exact functional form of the distribution functions.
4.9 PLASMA SHEATHS 215
Taking the logarithm of each side and solving for <l>w, we find
mi
$W (4.9.13)
me
kM = ! ln (4.9.14)
kT t me
which is about 2 for protons and 3 for heavier ions. Thus, the wall
potential is on the order of the thermal energy (divided by e) of the
particles in the plasma.
TlaXla 9 (4.9.15)
ax
f A common argument, for example, is to note that for 4> negative any positive
ion that passes through the edge of the sheath will continue and hit the wall. If
the velocity distribution at the sheath were a maxwellian with no drift, the flux
of ions would be [from Eq. (4.9.3)]
On the other hand, the electrons with insufficient velocity are repelled by the
potential and do not hit the wall; thus the number that do hit is still assumed to be
given by the flux estimate in Eq. (4.9.10), Te = n0(i<T/2meir)'12 exp (e4>u,/K/).
Such a calculation (because of the reduction in r») leads to a value for <t>„, which is
twice the value in Eq. (4.9.13). However, this is probably a lower bound for I\,
since the plasma at the edge of the sheath must have a net drift velocity toward
the wall, and for a drifting maxwellian I\ is easily found to be larger than I
Hence any modification to include this drift velocity would move the value for
|<l>u,| closer toward our original estimate (see, for example, Prob. 4.15).
216 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
and
dua dp a
Paf\^x) a
paUa~dx dx
d$> „ dna (4.9.16)
— naqa — — k1
dx dx
If we write out Eq. (4.9.16) explicitly for the ions and electrons,
we have
dui kT drii
— rriiUi
6 dx dx 7ii dx
due kT dne
= meue
dx
+ ne dx
(4.9.17)
dna
ua (4.9.18)
dx
Hence the ratio of the magnitudes of the two terms on the right side
of either part of Eq. (4.9.17) is
kT dna
na dx kT
(4.9.19)
dua maua2
'W^'OL^JOL j
dx
Thus the neglect of the two underlined terms in Eq. (4.9.17) requires
that the ratio of the thermal energy to the kinetic energy due to the
flow of plasma be large for the electrons and small for the ions, so that
t The analytic solution, however, gives <t> in terms of na, and one must use
numerical techniques to invert the equation and find na(<t>).
4.9 PLASMA SHEATHS 217
ne = n0 exp (4.9.22)
Note that this result is also what we should have obtained using Eq.
(4.9.9) for fe. (This is not surprising since the drift velocity is neg¬
lected in both cases.)
To find n, we integrate both the first part of Eq. (4.9.17) and
Eq. (4.9.15) to find
(4.9.25)
(4.9.26)
218 THE BOLTZMANN-VLASOV EQUATION FOR A PLASMA
Wall
Fig. 4.8 Plot of ion and electron number densities and of the potential in the
sheath region between a plasma and a wall (assuming = — 2kT/e).
Recall now that |e<t>| ranges from zero in the plasma to a value of order
kT at the wall and that [from Eq. (4.9.20)] we have already assumed
that miUo2 is larger than kT. Hence, if we restrict attention to the
region near the plasma edge of the sheath, we may assume that |e$| is
small compared to both kT and FmiUo2, and we can expand the terms
on the right side of Eq. (4.9.26) to find
$ = Aexlx (4.9.28)
(4.9.29)
[Recall that from Eq. (4.9.8) the expression on the right has the same
value for electrons and for ions.]
With this expression for u0 we can easily verify that the Bohm
criterion is satisfied, since we have (for e$w ~ —2kT)
(4.9.30)
Hence our assumption that |4>| grows exponentially rather than oscil¬
lates has some justification.
ue =
UqUo
< u0 exp
/ (4.9.32)
ne V -*t)
and
n0u0 ^
Ui = - > Uq (4.9.33)
Ui
kT kT 2e$w
> 2ir < 2tt exp 0.11 (4.9.34)
meue2 m%Ui2 kT
Problems
4.14Show that if we use Eq. (4.9.9) for/e and/,, the differential equa¬
tion for <t> reduces (in the region when e4> <3C kT) to
d24> _ <f>
'Uo
Ti n0 [exp ( — r2) + 7r1/2r(l + erf r)] r =
di
Use this solution together with the prior result, well away from the
wall,
3 kT
3 + ^+^ 4*2 —
4e
+ Ad + Ad2
(Note that 4>2 agrees closely with 4>i, even in the region from 0
to Ad.)
chapter five
THE BOLTZMANN
EQUATION FOR
A PLASMA
chapter five
(5.1.1)
dt + »< ir
% dXi + “.ir
dVi =
for the distribution function but we did not attempt to derive the
collision term (dfa/dt)c- Instead, we either neglected collisions and set
(dfa/dt)c = 0, or we replaced (dfa/dt)c by the simple Krook collision
model
mv + miVi
Co mo = m + mi (5.1.3)
m0
g = Vi — v (5.1.4)
5.1 TWO-BODY COLLISIONS AND THE BOLTZMANN COLLISION TERM 225
so that Cq, the center-of-mass velocity after the collision, clearly equals
c0. Next, if we let g be the reduced mass
mm i (5.1.6)
m0
g2 = g> 2 (5.1.9)
so that g', the magnitude of the relative velocity after the collision,
exactly equals g.
In view of the conservation of Co and g it is useful to describe
a collision in terms of a coordinate system where g lies in the 3 direction,
as shown in Fig. 5.1. (We shall refer to this as the gz coordinate sys¬
tem.) In this figure the angle x between g and g; is called the scattering
angle, while e defines the collision plane (the plane formed by g and g ).
The scattering angle x (or, equivalently, the vector k) is the only
quantity that depends on the details of the collision. I he technique
for evaluating x is detailed in Sec. A.2 of the Appendix. Here it is
sufficient to note that for force laws involving only the distance
between the particles x depends on three factors:
gx = gv = 0 gt = g
(5.1.10)
g'x = g sin x cos e g'y = g sin x sin e g'z = g cos x
(5.1.11)
(5.1.12)
Now we have shown that c'0 = co, and from Fig. 5.1 we can set
g' = g - ak (5.1.13)
If we use this expression for g in Eq. (5.1.11) and recall that Co = c0,
we find that
Inverse collisions
(5.1.19)
m(v) + rai(vi) —> m(v') +
b
228 THE BOLTZMANN EQUATION FOR A PLASMA
which shows that, for the given impact parameter b, the velocities of
the particles change from v and vi to and v'j. We now want to
prove that the inverse of Eq. (5.1.19) also holds, so that if the initial
velocities in the collision are v' and v( and the impact parameter is
still b, then the final velocities are v and vi, provided that the initial
positions of the particles are selected in such a way as to leave k
unchanged.
For this inverse collision the scattering angle is again x, since the
impact parameter, interparticle force law, and relative speed are all
the same as for a direct collision. Hence, depending on how we choose
the initial positions of the two particles, the vector g', which is now the
initial relative velocity, can swing either back to g or down to an angle
of 2x relative to g. The former case, which arises if we arrange the
particles initially so that k is unchanged, is called an inverse collision.
This is illustrated in Fig. 5.2, where we plot the path of particle mi
relative to particle m (located at the origin 0 in each plot) for a direct
and for an inverse collision.
If we denote the velocities after the inverse collision by a double
prime, we have
g" = g (5.1.20)
so that the inverse collision does indeed have the form m(v') +
mi(v[) —> m(v) + mi(vi).
b
To complete this discussion of two-body collisions we shall prove
the useful identity
dv'u dv’u
dvx dvu
If one uses Eq. (5.1.15) to calculate all the partial derivatives, one
finds, after a tedious but straightforward calculation, that J = — 1,
so that the identity is proved. However, the result \J\ = 1 can also
be inferred from Eqs. (5.1.15) and (5.1.18) without any calculation.
To see this recall that we can also set
d3v d3vi = \J'\ d3v' d3v[ = \J\ \J'\ d3v d3v 1 (5.1.25)
where J' is identical to ./ but with all the primed and unprimed quanti¬
ties interchanged. Now, from the symmetry of Eqs. (5.1.15) and
(5.1.18) one can easily prove that a partial derivative of any component
of v or vi with respect to any component of \' or Vj equals the same
partial derivative with primes and unprimes interchanged. [I'or
example, dvx/dv'ly = dv'Jdvly = 2(mx/m,)kykx.] Thus each element in
230 THE BOLTZMANN EQUATION FOR A PLASMA
H — // In / dh (5.2.1)
H = //(In / + 1) dh (5-2.2)
f Note that for problems with spatial gradients, Htot = Sf In / d3v dhr, so
that the H used here is Htot Per un't volume.
232 THE BOLTZMANN EQUATION FOR A PLASMA
always have opposite signs (unless both are zero). Hence their product
is negative semidefinite, and we have
H < 0 (5.2.9)
fh = f'fi (5-2.10)
This result shows that when one uses the Boltzmann collision term, H
always decreases (or, conversely, the entropy always increases) with
time until ffi =; /'/(. When this occurs, there is no further change
in the system, since, from Eq. (5.2.3), / = 0.
Problems
n7e_cW<V Cy v u')'
= T3'2ayz , 2 KTy
my
and y = a or j3. Show that this difference goes to zero if and only
if ua = up and Ta — T0.
5.3 (a) Show that for a simple gas, if / = C(x,t)e~0<-I’t)v2, then /
satisfies the Boltzmann equation if and only if /3 is constant
and C(x,t) = Cie~2U(z)fflm, where U(x) is the potential of any
5.2 THE H THEOREM 233
The H theorem predicts that the entropy always increases with time.
While this agrees with the second law of thermodynamics, the result is
nevertheless not compatible with the fundamental laws of mechanics,
which are reversible. Thus in a system of N interacting particles, if
we reverse the velocity of each particle, the laws of mechanics require
each particle in the system to retrace its former path. To see how the
H theorem leads to a paradox, consider the behavior of the following
two systems of N particles numbered from i = 1 to N. Let system I
be a nonequilibrium gas where the initial conditions for each particle
are given by x^O), v/(0) and where the initial value for H is H (0).
[Assume also that the Velocity distribution at t = 0 satisfies the con¬
dition /r(v) = /r( — v).] For system II, let the initial conditions be
defined by
In other words, each particle in the second system starts off initially
with the position and speed of the corresponding particle in system I
at time r but with the velocity reversed. Hence, from the laws of
mechanics,
I!
H H
System I System II
T T
and
Xf bf d3c = 0 (5.3.6)
From this result it is evident that Uif bf d3c also equals zero; hence,
if we multiply each component of Eq. (5.3.3) by a Lagrange multiplier
yi/m, we have
or
Finally, from the equation for the variation of the energy density
[Eq. (5.3.4)], we have, after multiplication by the Lagrange multiplier
2/3/m,
However, since we have already shown in Eqs. (5.3.6) and (5.3.8) that
the second and third terms here integrate identically to zero, we have
or
Since the integrand here is positive definite, this means that yk must
be zero. (Alternatively, one might set C = 0 or d = 00, but either
of these assumptions makes the distribution function equal zero.)
The final two constants are obtained from the definitions of the
number density and temperature
3/2
d3c = C (5.3.16)
n
238 THE BOLTZMANN EQUATION FOR A PLASMA
and
_dn _ 3C /tr\3/2
(5.3.17)
d(3 _ 2/3 \p)
Solving these two equations for & and C, one easily finds
m
/3 = or2 C = (5.3.18)
27r tr3/2a3
Tie c2 la2
(5.3.19)
7r3/2a3
Problem
5.4 (a) Calculate Htot = // In / dzv dzx for a gas with a maxwellian
distribution. Show that if we let S = — kHtot, then S satisfies
the thermodynamic relations
(S)„ = f and =I
with iV the total number of particles, V the volume, and
E — §NkT the total energy.
(b) Prove that if / = [w(x)/7r3/2a3] exp ( —i>2/a2) and a is assumed
constant, then f/tot is a minimum when w(x) is constant.
(Assume gas has N molecules in a fixed volume V.) Hint:
Use a Lagrange multiplier technique.
From our discussion at the end of the last section it is evident that the
maxwellian velocity distribution, besides being the equilibrium solu¬
tion to the Boltzmann equation, is also the most probable distribution
one would have if the macroscopic state were specified by n, u, and T.
For nonequilibrium problems one can still perform a variational cal¬
culation of the type just shown, to find the most probable distribution
(subject to the constraints provided by a set of macroscopic param-
5.3 MAXIMUM ENTROPY AND THE MAXWELLIAN DISTRIBUTION FUNCTION 239
H = lH Ha I fa In fa dh (5.3.20)
<a
Next, since n«, u„, and Ta must all remain fixed as fa is varied, we can
set <5n«, 5(c„), and 5(3naxT«/m«) all equal to zero, to find
where each gas has a maxwellian velocity distribution but with its
own temperature and mean velocity. Note again that this is not an
equilibrium distribution (unless the temperatures and mean velocities
of all species are equal), since, as shown in Prob. 5.2, the collision teim
for such a mixture does not vanish. However, it is the most probable
distribution with these constraints.
Problem
5.5 Find the most probable value for / for a one-component non¬
equilibrium gas when the known macroscopic parameters aie the
240 THE BOLTZMANN EQUATION FOR A PLASMA
(5.4.1)
(5.4.2)
p
t In fact, since the truncated solution for / is generally not an exact solution
to the Boltzmann equation, the results obtained are not independent of the choice
of the transport equations to be used. While this problem has not been thoroughly
investigated, it appears that the best results are obtained when the least compli¬
cated Q’s are used to obtain transport equations.
5.4 LOWEST-ORDER TRANSPORT EQUATIONS FOR A GAS MIXTURE 241
relatively simple example of the method, where we use just the first
term in Eq. (5.4.2) for each species of particles in the plasma. Although
this is an exceedingly crude level of approximation for/a, we shall see
that it leads to a full set of mass, momentum, and energy equations
(including collisional interaction terms) for each species—much like
the equations used in Chap. 3 but with the collision terms for momen¬
tum and energy transfer given explicitly (instead of in terms of some
empirical collision frequencies).
For this example we set
nae~Cai,aa'1 2 kTc
u = ca = v — u aj (5.4.3)
7r3/2aa3 ma
where na, u«, and Ta may all be functions of position and time. (Recall
that this would be our most probable form for fa, if we “know” values
for na, ua, and Ta.) For this case we have five functions of position
and time for each species (since ua is a vector); hence we can use the
five transport equations for the mass, momentum, and energy of each
species that are given in Eqs. (4.2.12), (4.2.16), and (4.2.29) but which
we rewrite here for convenience.
Pa + V • (paUa) - 0 (5.4.4)
Duak (5.4.5)
Pa(&k)a ^ Paik &Qak
pa~i)r
D
PaPa
—5/3
V • qa (5.4.6)
Dt
where
(5.4.7)
(5.4.8)
(a)a = — (E + u„ X B) (5.4.9)
m
Next, recall that paik = pa(CaiCak), so that, using Eq. (5.4.3) for /«,
This integral is zero when i ^ k, since the integrations over cai- and
Cak are of the form
J dx (5.4.11)
Pa^a ~ ^
Paik = ^ Pa^ik (5.4.14)
so that the traceless pressure tensor, for this case, is zero. Similarly,
for the heat-flow vector, we have, by definition, qak = -ipa(ca2cak), so
5.4 LOWEST-ORDER TRANSPORT EQUATIONS FOR A GAS MIXTURE 243
= 2^k? f d’e° 0
=
(5.4.16)
since the integral over cak again has an odd integrand. In summary,
we see that when fa is a maxwellian velocity distribution, the traceless
pressure tensor and heat-flow vectors are identically zero, so that the
transport equations given in Eqs. (5.4.4) to (5.4.6) reduce to
Pa ^ ‘ (paUa) — 0
(5.4.17)
Z) (5.4.19)
PaPa~bli = fPa 5/3 &QaO
Dt
(5.4.20)
8Q = J / (J'J'pi - fMQgbdb dt dh dh'
e
where ai, a2, and a3 are unit vectors in the x, y, and z directions in the
gz system. Hence, to find the fcth component of g and g' in some
external coordinate system we take the dot product with a* to have
Qk = <7«8*
(5.4.24)
9k = {/[sin x{oL\k cos e + a2k sin e) + a^k cos xl
and
(5.4.34)
0
and
8pga0 r«
Vap = e~x2/Sap(1)xb dx (5.4.36)
3m07r1/2 Jo
law are called Maxwell molecules.) For this case, from Eq. (5.4.31),
the cross section is proportional to g~l, so that
with
Now, from the definitions of na, rip, ua, and up, we have for any fa
and fpi,
but with the collision frequency given explicitly now by the expression
while
(5.5.8)
g = C0! — C„ + (up — U«)
248 THE BOLTZMANN EQUATION FOR A PLASMA
where A is the sum of all terms proportional to cai or Cpu and in our
usual notation for the square of a vector
Hence, if we replace Cafj by Eq. (5.5.3) and use ra/3 as defined in Eq.
(5.5.6), we have
fiQak (5.5.13)
L ir3da3dp3
e
g dp2ca + aa2c?i
x = y =
do CloClaClp
(5.5.14)
11/3 “ Ua
£ = - do2 = da2 + dp2
a0
From these definitions one easily finds that the factor in the exponential
= (x - e)2 + ya (5-5-15)
da2 dp2
When we substitute these results into Eq. (5.5.13) and perform the
resulting integration over d3y (which is now straightforward), we find
(5.5.17)
SQak ^ HQaPIc
(3
where
of each component of x.] For the next term in the expansion we have
rv T
Hence, if we set e = 0 and apply the identity, which again holds for
any F(x) that depends on the magnitude of xf
we have
87T
e “’Sa/s^x6 dx (5.5.23)
3"
Applying this result in Eqs. (5.5.17) to (5.5.19) for 8Qak, we find that
8Qak can again be written (to this lowest nonvanishing order in e) in
f Note that in the last step of this identity we merely switch to spherical
coordinates, where d3x = x2 sin 6 dx dd d4>, and then integrate over 9 and <j>.
5.5 CALCULATION OF COLLISION FREQUENCIES 251
the form
Sng/xQp (5.5.25)
Vaf! = e-^Sa^x5 dx
3ma7r1/2 Jo
where p is the inverse power in the force law and n = — 4/(p — 1).
For A*(p) to be finite, p must be greaterf than 2; hence, as shown in
Fig. 5.5, n lies between -4, for coulomb forces, and 0, for hard spheres
(where p —* °o). Using this expression for Sa(3(1), and noting that
— (l/2)n
87r1/2P3d.i(p)aon+ir(3 + vn)
Va& =
3m0 © (5.5.28)
f For coulomb forces, with p = 2, the divergence in A tip) is avoided (as shown
in Sec. A.4 of the Appendix) by considering only impact parameters less than the
Debye length \D. The justification for this is based upon our discussion of the
Debye potential in Sec. 4.8, where we noted that the coulomb potential 4>f between
two particles in a plasma is effectively reduced by the factor exp ( —r/Xn). Since
the full Debye potential is hard to use in calculating Ah one still uses 4>c but just
for b < Xz>.
252 THE BOLTZMANN EQUATION FOR A PLASMA
TZI2
A = 9.0A"d = 1.23 X 107 --/2 (5.5.29)
and <x is the sum of the radii of the colliding particles (for ven this is
essentially the classical radius of the neutral molecule). If we sub¬
stitute these values for A j and n into Eq. (5.5.28), we find
and
TABLE 5.1
Coulomb 47r(e2/47T€ong2)2 In A -4 2 In A
Hard spherej 7rcr2 0 cr2/2
1/2
to
(2k)1/2
meTA112 (5.5.34)
— ^ 2.8
V en mlTe)
(5.5.35)
C Qi 'llai
a0
Vi + Gi(x)
where
2k (5.5.36)
Gi(x) {Tf> - Ta){xi - et) +^a0€,-
254 THE BOLTZMANN EQUATION FOR A PLASMA
we find
aaa0
8Qa0 = £ Mny°- / exp [-(x - e)2 - t/]Sap (i) Vi
a0
B
with
Comparing this result with Eq. (5.5.18) for 8Qapk, we see that
Problems
In the last section we analyzed 8Qa in the limit when e —> 0, and found
that for all inverse-power interparticle force laws
with vap given, for arbitrary force laws, by Eq. (5.5.25) or (5.5.28).
Hence, for small e it is clear that 5Qapk increases linearly with the
difference in the flow velocity. To illustrate the significance of this
fact consider a fully ionized gas in the presence of a constant electric
field. Assuming no spatial gradients and that B is negligibly small,
we have from Eq. (5.4.IS)
n, = E Uj) (5.6.2)
me
so that, after a time large compared to v~\ the difference in the flow
velocities becomes constant and equals eR/^v. At equilibiium the
current density, as might be expected from our analysis of the ( (in¬
ductivity in Sec. 2.2, is given by
(5.6.8)
with
nanffe4 In A
(5.6.9)
47r5/2eoVa02
system
and
where the aik are direction cosines of the fth axis in this collision-integral
coordinate system relative to the /cth direction in the external coordi¬
nate system (atk = a, • a*). Using Eq. (5.6.11) to handle the 4>
integration, we find
IT il2Ctke- (5.6.13)
foQaflk es2s(l + erf s) ds
The first term in this final integral is zero, since the integrand is odd;
while the second term (which has an even integrand) is easily evaluated
by integrating by parts. In the formula fU dV = UV| - J1 dU
we letj
j To take the derivative of the error function recall the formula for the
derivative of an integral given in Sec. 4.6:
Fig. 5.7 Plot of \p versus e for coulomb forces. For comparison, the result
is compared with the straight line = e, which holds for Maxwell molecules
and is also the small e approximation for all force laws.
to find
From Eqs. (5.5.28) and (5.6.9) and Table 5.1 we can rewrite 8Qap,
where 8Qag = [^ (SQ^fr)2]1/2, in the formj (useful for computation)
k
Electron-runaway effect
Using the exact expression for 5Qei and 8Qie, suppose we reexamine the
interaction of the plasma with a steady electric field. The momentum
equations for the electrons and ions [Eq. (5.4.18)] with all spatial
jp C
&Qeik i (5.6.18)
Uek —--Uk ~I-
me Pe
6 77T . &Qiek
(5.6.19)
Uik = — -C/k H-
mi pi
Hence, by subtracting the first of these from the second and using
Eqs. (5.6.16) and (5.6.17) we find
eE a0d/'(€)rA
(5.6.20)
A - u il
A
with V = Vei + Vie. This equation has two classes of solution, depend¬
ing on the magnitude of E:
e (5.6.21)
Ec = 0.193 In A
47re0Az32
Since In A is of order 10, this shows that Ec is on the order of the field
strength (in a vacuum) one Debye length away from a charge. In
addition, since A and XD2 are both proportional to temperature, the
temperature dependence of Ec is of the form
Ec * 71-1 In T (5.6.22)
Hence, runaway can occur for arbitrarily weak fields if the temperature
becomes sufficiently high.
260 THE BOLTZMANN EQUATION FOR A PLASMA
Problem
5.8 Prove that for an arbitrary inverse-power interparticle force law
Hence, for large e, <5Qa/3 —> 0 for all force laws with n < —2 or
p < 3; while for the other (shorter-range) force laws 8Qap—> °°
for large e, so that the runaway effect cannot occur.
The analysis in the last three sections is all based upon the assump¬
tion that each species of particles in the plasma has a maxwellian
velocity distribution, so that/a is specified by just the five “moments,”
na, ua, and Ta. With this assumption we were able to derive transport
equations similar to those developed in Chap. 3 on macroscopic grounds.
However, the phenomenological collision frequencies used in Chap. 3
could be computed from a knowledge of the interparticle force laws,
and there were some effects (most notably, the electron-runaway effect)
that one would not suspect from the macroscopic theory. All these
results, however, still omit such phenomena as heat conduction and
viscous-damping effects in the plasma.
To analyze these factors we must extend the analysis in some way.
However, to reduce the algebraic detail involved in such an extension,
we shall, in this section, restrict our attention once again to an electron
gas in a stationary background of heavy particles which has just
enough uniformly distributed positive charges to make the plasma
macroscopically neutral in equilibrium. In addition, rather than
using the full Boltzmann collision term, we again use the Krook colli¬
sion model, (df/dt)c = — v(J — /0), whose use greatly simplifies the
collision integral.!
f For this case we must assume that v is the electron self-collision frequency,
since, as noted in Sec. 4.3, any attempt to interpret v as an electron-heavy particle
collision frequency leads to large errors in at least some of the collision terms.
Nevertheless, the results we shall obtain for the electron pressure tensor (or vis¬
cosity) and heat-flow vector (or thermal conductivity) in this section and the next
can be put into remarkable agreement with those obtained by exact calculations
for a weakly ionized gas if we do interpret v as the electron—heavy particle collision
frequency.
5.7 THE THIRTEEN-MOMENT EQUATIONS FOR AN ELECTRON GAS 261
-c^/a2
ne (5.7.2)
/o
7r3/2a3
and where Aif C<, and B{j ( = £*), like n, u, and T, are functions of
position and time but not of v. (Hence we shall refer to them as
constants.) The reasons for expanding in a power series in c are two¬
fold: (1) the integrals involving / are all manageable, and (2) the coeffi¬
cients Ai, Bij, and C; all have a real physical meaning.
To show this we need only compute n, (ck), pkh, and qk in terms of
the constants. To simplify the calculation we note, first, the following:
TABLE OF INTEGRALS
The integrals in this table can all be verified very quickly by using
the following scheme. First, note that for /o given by Fq. (5.7.2)
dfo (5.7.10)
dCi
Hence, for any integral involving the product foCi we can first use this
identity and then integrate by parts to find [assuming/0F(c) vanishes
at Ci = ± °° ]
I f0CiF(c) d3c
(5.7.11)
From these results one can easily, verify all the remaining integrals
with very little effort.
Through the use of this set of integrals we can now calculate n, (ck),
Pkh, and qk directly from their definitions in Sec. 4.1. We have, for
/ given by Eq. (5.7.1),
This result shows at once that Bn, the trace of B, is zero. Next, we
calculate p(ck), to find
C„ = - ^ (5-7.16)
5 a1
Thus far we have not yet related the coefficients to any physical
quantities, but this fault is remedied if we calculate pki and qk. From
the definition of the pressure tensor, we have from Eqs. (5.7.1), (5.7.3),
and (5.7.5)
(Note that in the last step we have used the fact that Bydij = Bu = 0.)
Comparing this result with Eq. (4.2.20) for the traceless pressure
tensor, it is evident that
P kl (5.7.18)
Bki
pa2
qk = im\fc2ck d3c
5 pa2 (5.7.19)
IT (Ak + ^o-2Ck)
Eliminating Ck through the use of Eq. (5.7.16), we find
A_2 qk (5.7.20)
Ak pa2
Pn P13
P = P12 P 23 (5.7.21)
P13 — P11 — P 22
Problems
Viscous-stress equation
5? + • u = ° (5.7.22)
DUk C rp i / w i 1 dpik
n, = [Pk + (u X B)fc (5.7.23)
Dt m p dXi
First, for Q = mckCj, we find for each term on the left side of this
equation
dpkj
Jtn{Q) dt
~n(Q) 0
— n (a,
dQ\ ■ p(djCk + cqe,) — — Mjk
dVi/
M ii — 2ojcPi2 M 22 = — 2o)CP 12 M 33 = 0
When Eqs. (5.7.30) and (5.7.32) are used, the viscous-stress equation
(5.7.29) involves just the 13 moments used in the approximation for/.
dqk
hn{Q) dt
~n(Q) Uj(Pjk T ^pSjk)
— n (a.
dQ\ p(Uj(CjCfc T ^C25yfc))
7 dvJ
8Q — vqk
When adding these terms together we note that the three terms pro¬
portional to pjk + |p8jk all reduce, because of the momentum equation
(5.7.23), to
Problems
5.11 Verify that one obtains the energy equation by letting/ = k (and
summing over k) in the viscous-stress equation (5.7.29).
5.12 Calculate the collision integral for the heat-flow equation when
the Boltzmann collision term is used and the Maxwell interparticle
force law is assumed, and show that, as here, the term is propor¬
tional to qk-
Pkj c. + £ + 3 +
V VL
(5.8.1)
qk Da + ^+^ +
V V
268 THE BOLTZMANN EQUATION FOR A PLASMA
(5.8.3)
(5.8.4)
Dp _ 5 p Dp
-fpV • u (5.8.5)
Tri ~ 3~pDt
dUj
Pkj = ^ih3v u (5.8.6)
dxk
5p d p _ dT 5 Kp
(5.8.7)
2v dxk p °dxk 2m v
The coefficients r/o and X0 which enter here are the viscosity and the
thermal conductivity of this simple electron gas. The ratio of X0 to 170
is easily seen to be 5k/2to. Had we performed this calculation using the
Boltzmann collision term (just for self-collisions), this ratio would
instead have been 15/c/4m, so that the error involved in using the sim¬
ple Ivrook model (where, as noted in Sec. 4.3, the relaxation time for
each quantity is exactly r_1) is only a factor of 1.5.
If we substitute the expressions in Eqs. (5.8.6) and (5.8.7) for
Pkj and qk into the momentum and energy equations (5.7.23) and
5.8 CALCULATION OF TRANSPORT COEFFICIENTS 269
Duk e ,„ , „ -\ 1 dp 170
-7T7 =-[Bk + (u X B)fc---1- V2Ujfc + ~(V'u)
Dt m p dxk p 3 ax*
(5.8.8)
,5/3 IL
Dt
VP
-5/3 _
•fXo
v 1 + ^ dxi
(5.8.9)
Co = D0 = 0 (5.8.11)
However, to next order the results are quite different. First, in the
heat-flow equation [Eq. (5.7.35)] for B = Bat, we now have, in place
of Eq. (5.8.4),
(5.8.12)
|p — - - wc(q X at)k = -D1
2 dxk p
q = -A-VT (5.8.14)
with
Ax Ah 0
A = — A// Ax 0 (5.8.15)
0 0 Ao
and
\ _ A0r2 A0cocr
= (5.8.16)
Ax ~~ W„2 + C
This residt shows that for a fixed temperature gradient the heat
flow in the direction of B is unchanged. However, the heat flow is
reduced in the directions perpendicular to B and, in addition, there is a
Hall heat flow in the direction perpendicular to both B and VT.
Calculations of X± and \H as a function of magnetic-field strength are
shown in Fig. 5.8, where it can be seen that both fall off fairly rapidly
with increasing values of o>c.
p(-^-“+S+S)_Mh= (5-8'17)
with Mkj given in Eq. (5.7.32). If we divide each term by v and recall
that 770 = p/v and that, to this level of approximation, P kj = Ci/v,
we have
> (5.8.18)
II
1
, „s _
fkj - ?hjV • u
dUj
^ ■ d-ff± (5.8.19)
dxj
P 33 = 11of33
and
II
1—1
Pn -2 -CP12
V
S* Pn + P„ - ^ Pn = Wi* (5-8.21)
V V
2 — P12 + P 22 = 770/22
V
The set of equations involving P 1.3, P23, and P33 is again exactly
like the current-flow equations with Pk3 replacing Jk, fk3 replacing Ek,
and 770 replacing cr0. Hence, as in Eq. (2.2.7), we have
with
The second set of equations (involving Pn, P12, and P22) is easily
solved to find (using the identity fn + f22 = —/33)
In the limit when coc/v approaches zero these results reduce again
to the form
-I 0 0
P — ^0/33 0 -i 0 (5.8.26)
0 0 1
From these results it is clear that viscous effects with a strong mag¬
netic field are generally quite different from those found using the
usual Navier-Stokes equation.
Problems
for the stress and heat-flow equations. Although all the integrations
can be carried out for many inverse-power interparticle force laws,
the results are quite complicated and have not yet been reported in a
form which is easy to interpret physically. Here we note only that
in the stress equation, for example, where the Krook model leads to
-vPak] (5.8.28)
a simple expression for the dc conductivity and, again, using the five-
moment approximation for the distribution function in Sec. 5.6 to
study the electron-runaway effect in a fully ionized plasma. Now we
shall apply the Boltzmann equation to a very weakly ionized gas, so
that the collision term for the electrons is dominated by the electron-
neutral collisions. (This is generally called a Lorentz gas.) The
questions we pose are the following: (1) what is the velocity distribu¬
tion of the electrons when there is an applied uniform electric field,
and (2) what is the current flow (or conductivity) in the plasma?
For simplicity we assume that there is no magnetic field present,
that the plasma is completely uniform (so that there are no spatial
gradients), and that a steady state has been attained (so that there
are also no time-derivative terms). With these assumptions the
Boltzmann equation for /(v), the velocity distribution function for
the electrons, reduces to
(5.9.1)
(5.9.2)
In this collision term F(V) is the velocity distribution function for the
neutral molecules, and, as usual, f and F' are the distribution functions
evaluated after the collision. It is assumed that F(\) is known, and
in practice we take it, later, to be a function of just the magnitude of V.
Although Eq. (5.9.1) appears to be relatively simple, it is still
too complicated to handle exactly. Hence we shall introduce another
expansion procedure for/. We let
where /0, fi, fa, etc., are all functions of just the magnitude of v, so
that, for any one of them, say /p,
e_ df0 d fi
m v
Vk -J-
dv
+ ViVk ~J - + ft
dv v
(5.9.5)
J J Vi sin 6 dd d<f> = 0
(5.9.6)
JJ ViVk sin 6 dd d 4> = A
When these identities are applied to the integration over 0 and <t>,
Eq. (5.9.5) reduces to
x for v.
276 THE BOLTZMANN EQUATION FOR A PLASMA
1 rf 2/ _ * A (5.9.8)
3v dv 1 k dv dv
4ireEk d_ 2,
sin 6 d9 d<fi (5.9.9)
3mv2 dv 1 k //(
The collision term on the right side of this equation (like most
collision integrals) is difficult to evaluate. Therefore, before attempt¬
ing to calculate it we shall first obtain a set of three additional differen¬
tial equations, so that we have four equations to determine /0 and the
three/i functions. To obtain these other equations we simply multi¬
ply Eq. (5.9.5) by Vj sin 9 d9 cl(j> and integrate over 9 and <f>. From
the identities in Eq. (5.9.6) we find
(5.9.10)
where j can take on the values 1,2, and 3 (or x, y, and z). We now
have enough equations to solve for /0 and the /,- functions. However,
before proceeding further we must evaluate the collision terms on the
right side of each equation.
F gb db de d3V (5.9.11)
To evaluate this term we note first that the mass of the neutral molecule
M greatly exceeds the electron mass m and that
(5.9.12)
where by definition
(5.9.18)
\(V) = (UnSen^)-1
is the mean free path for the electrons. If we apply this result to the
collision terms on the right-hand side of Eqs. (5.9.9) and (5.9.10), we
find [after multiplying by -3y/47r and using the identities in Eq.
(5.9.6)] that
(5-9-19)
and
dfo (5.9.20)
dv s//"J («).““9=
278 THE BOLTZMANN EQUATION FOR A PLASMA
From the second equation here we see that each of the f}(v) functions is
simply related to f0(v) by the condition
However, because of the null result in the first collision term, we shall
have to evaluate (df/dt)c to one higher order (where some account is
taken of the finite mass and velocity of the neutral molecules) in order
to complete the problem.
Jk = - ~ JQ vzfk dv (5.9.23)
Jk — <to Ek (5.9.24)
(5.9.25)
1 V
X (5.9.26)
V
(5.9.27)
However, from the definition of the number density and the identities
given in Eq. (5.9.6), we find that for / = /0 + (Vi/v)fi
Hence, for this case the conductivity reduces to the familiar result
<r0 = —2 (5.9.29)
mv
Solution for /0
For other force laws it is not possible to calculate the conductivity in
Eq. (5.9.25) unless one can find /0. However, in order to find /0 we
need to use a more accurate expression for (df/dt)c, so that the collision
term in Eq. (5.9.19) does not entirely vanish. This next-order calcula¬
tion of the collision term is carried out by a rather straightforward
expansion in powers of m/M [assuming also, from Eq. (5.9.12), that V2
is of order (m/M)v2\ in Sec. A.6 of the Appendix. Here we shall merely
state the result that to lowest nonvanishing order in the collision term
we find, in place of Eq. (5.9.19),
with A2 = 2\iTn/M.
The determination of /0 is now straightforward. We first use Eq.
(5.9.21) to eliminate/* in the term on the left and let y = eE/m, to have
dfo mv dv (5.9.33)
77 ~~ ~ (MX2y2/3v2) + kTh
so thatf
v' dv'
(5.9.34)
fo - C exp
MX2y2/3v'2 + kTu )
This expression for /0 is essentially our final result. The integration
in the exponential can be performed but only after the interparticle
force law (or A) is specified. Thereafter, the integration constant C can
be expressed in terms of the number density by using Eq. (5.9.28).
Finally, from Eqs. (5.9.25) and (5.9.21) we can calculate the conduc¬
tivity ao and the various /*■ distribution functions.
v = 2Tr(A/m)1/2n„Ai(5) (5.9.35)
Using this expression for A in Eq. (5.9.34) and evaluating C by the use
of Eq. (5.9.28), we easily find that/0 is a maxwellian velocity distribu¬
tion of the form
nce-!,2/a*2
(5.9.36)
7r3/2a*3
but with
My2\
a *2 + 3r2 )
(5.9.37)
t Note that this result differs by a factor of 2 in the term with y2 from the dc
limit obtained for /0 when there is an applied oscillating E field. This is because
for an oscillating field (E2) = ^E2, whereas here we require no such average over
time [see, for example, H. Margenau, Phys. Rev., 69:508 (1946)].
5.9 DISTRIBUTION FUNCTION FOR A STRONG £ FIELD 281
y* My2
Tn + (5.9.38)
3v2k
fo = fo(Tn = 0) + _q Tn + • • • (5.9.40)
U_mXy (5.9.42)
fo xTn + (M\2y2/3v2)
This result suggests, first, that the expansion procedure converges more
rapidly for the slow electrons than for the fast ones. In addition, the
282 THE BOLTZMANN EQUATION FOR A PLASMA
Problems
5.15 In the limit when Tn is negligibly small calculate a0, u (the average
velocity in the direction of E) and (v2) for an arbitrary inverse-
power interparticle force law. Show that, for a strong E field,
(v2) » w2, so that, to a good approximation,
m(c2) _ m(v2)
3k — 3k
(5.10.1)
fQ(f),d,” = ~ IQ*•(&*>-)
+ i J Q d3v (^dv9dl, Ja(AVi — • • • (5.10.4)
where
and
Since Eq. (5.10.4) must hold for any function Q(v), we conclude
that
(5-ia7)
This is the Fokker-Planck collision term, and the quantities (Avt)av and
(AVi AVj)av are known respectively as the Fokker-Planck coefficients of
dynamical friction and of diffusion (or diffusion in velocity).
While the expansion procedure used to obtain the Fokker-Planck
collision term can, in principle, be extended to any number of terms, in
284 THE BOLTZMANN EQUATION FOR A PLASMA
practice only the two terms shown in Eq. (5.10.7) are ever used. For
this reason it is a reasonable approximation to (dfa/dt)c only when Av,
the change in velocity due to a collision, is small for most collisions.
While it is hard to state this condition very precisely, it is generally
assumed that the Fokker-Planck term is a good approximation for long-
range forces, such as the coulomb interaction between charged particles,
and a poor approximation for short-range forces, such as the interaction
of hard spheres.
(5.10.10)
(5.10.11)
where an and an are the dot products, a* • a, and a„ • a,-. (Note the
gz coordinate system repeated here in Fig. 5.10.) Hence, for the pro-
5.10 THE FOKKER-PLANCK EQUATION 285
Our result is now in a form that is easy to integrate over all impact
parameters. Recall that we have previously [in Eq. (5.4.30)] intro-
Hence, if we use this definition and the result obtained in Eq. (5.10.15),
we find, for our collision integral,
(5.10.18)
(5.10.19)
with the coefficients (Ar,)av and (Ai>,- AVj)av given, respectively, by Eqs.
(5.10.10) and (5.10.18). These general expressions are so formidable
in appearance that one might well question whether the Fokker-Planck
collision model is any simpler to use than the full Boltzmann equation.
However in many specific cases, particularly those involving charged-
particle interactions, the final collision term can be reduced to a rela¬
tively simple result.
n0C\Vj
(A Vi)a
v3
(5.10.23)
with
■ d2f
+ if (t~ (Al>i A^a
dVi \OVj
1
2 dVi dVj
(AAt^')av (5.10.25)
Each term on the right side of this equation reduces to a simple expres¬
sion when (Avj)av and (Aw,- Awy).v are given by Eq. (5.10.23). However
some algebra is required before we can obtain our final result. We
note first the following useful identities:
A = _ A hi = ill (5.10.26)
dViV3 dVj v3 dvj v v3
f As usual this assumption would break down in those rare cases when Ti/rry
is comparable to TJme.
288 THE BOLTZMANN EQUATION FOR A PLASMA
From these identities and Eqs. (5.10.23) and (5.10.24) we find that
while
2 n0Vi
— (AVi Avj)a (Ci - C2) (5.10.28)
dv.
CzVj df d2/
n0
v3 dVi
+ dVi dvj
(5.10.29)
where
This residt is now in a form which can be used easily for many problems.
For other types of coulomb interactions, the Fokker-Planck col¬
lision term also reduces to a relatively simple form even if we cannot
assume that the j8 particles are heavy and stationary. To see this we
note first that since g = vx — v, if we have any function F(g), then
dF _ _ dF
(5.10.31)
dVi dgi
_d_ d2
( A?^i)av (AVi AVj)av = 0 (5.10.32)
dVi di>i dVj
and
mpC 1
(Ci — C 2) +
m0
Problems
5.17 Investigate the Fokker-Planck coefficients for Maxwell molecules
(where the results are independent of fp 1) and compare the results
with Eq. (5.10.23).
5.18 Evaluate the full expression for (dfa/dt)c for electron-electron
collisions when/01 is assumed to be maxwellian. Hint: You may
refer back to Eq. (5.6.8) to handle the integrals.
chapter six
THE BBGKY
EQUATIONS
FOR A PLASMA
chapter six
In the last two chapters we have been using a statistical method for
describing a system of particles in terms ol the particle distribution
functions fa(x,\,t), where fa(x,v,t) dzx d3v is the number of particles
of type a in the volume element d3x with velocities in the range d3v at
time t. Although a knowledge of fa enables one to calculate macro¬
scopic variables such as the pressure, temperature, mean velocity, and
so forth, we have seen that some difficulties remain in this approach.
Mainly these difficulties arise in the attempt to include particle inter¬
actions or collisions in the partial differential equation for fa. The
various approaches that we have seen for handling the collision term
range from the collisionless BV equation and the relatively simple
Krook collision model to the full Boltzmann collision term and the
Fokker-Planck approximation to it. In all these, however, funda¬
mental assumptions are used (such as those delineated in our deriva¬
tion of the Boltzmann collision term in Sec. A.3 of the Appendix)
whose limitations are not easy to assess.
For this reason, in this chapter we shall introduce a different
approach (developed independently by Bogoliubov, Born and Green,
Kirkwood, and Yvonf), in which, initially, the particle interactions
are handled exactly. Needless to say, this exact treatment of the
interactions leads, as we shall see, to difficulties of other kinds that
make it virtually impossible to apply the approach to most of the
practical problems we have solved using the Boltzmann equation.
However, the method has been applied to some problems and can also
be used as a means of clarifying the approximations involved in using
the BV equation or any of the more detailed collision models. Before
embarking on our analysis of the BBGIvY equations we should note
that almost any analysis using the BBGIvY equations becomes
extremely involved. Hence, in this chapter we shall cover only the
derivation of the equations themselves and some of the general pro-
f It is for this reason that the resulting equations are called the BBGKY
equations.
6.1 THE LIOUVILLE EQUATION 293
cedures that are used in working with them. The attempt to apply
the equations to actual problems represents an area of current research
activity which we shall not attempt to cover here.j
N
/<"(!,2, . . . N,t) n d3xad3va (6.1.2)
a=l
side of this equation and the same factor (evaluated at t + A<) on the left. How¬
ever, as in our derivation of the Boltzmann equation in Chap. 4, these differential
elements differ only by terms of order (A<)2 and drop out in the limit as At —> 0.
294 THE BBGKY EQUATIONS FOR A PLASMA
If we expand the term on the left about its value when At = 0, we have
Thus far we have used only f(N), the joint probability function for all
the particles in the system. If we want to find the probability that
some smaller group of particles has a particular set of positions and
velocities, all we have to do is integrate /(A0 over all possible positions
and velocities for all the other particles. Thus if we let
then plainly/(n) is the joint probability function for the first n particles
irrespective of the positions and velocities of all the remaining particles.
In particular, the probability function for just the first particle is
Problem
6.1 Write out the Jacobian matrix for the transformation from xa(t),
xa(t) to xa(t + At), xa(t + At), where a runs from 1 to N. You may
assume that aa (the acceleration on particle a) depends on the
positions of all particles. Show that the determinant has no
terms proportional to Ah
N
dfW df(1)
Ti = [] d3xa d3va (6.2.1)
/ dt
a=2
dt
since, as the limits are independent of time, d/dt can be taken outside
the integral. Next, consider the integration of the second term in the
Liouville equation. For simplicity we separate out the first term in
the summation over a to have
To =
J
f vu
V d-P n
dX" 8 = 2
d'x, dh» + ff
a=2 J
va% IP n
dXal 0.2
d% (6.2.2)
dfw +»
dxai = fw no summation over i (6.2.3)
/- “ dXai %r--
T dfw 11
n d+pd+fj
js df(1) (6.2.4)
1 2 VI i+r- = Vii
I dx u
Xlt p = 2
296 THE BBGKY EQUATIONS FOR A PLASMA
aa = A„ + ^ aa7 (6.2.5)
7=1
with Aa the acceleration on the ath particle due to external forces and
aa7 the acceleration on the ath particle due to its interaction with the
7th particle. The prime on the summation over 7 is to note that the
term 7 = a is excluded from the sum, since the particle does not
interact with itself.
For the term involving Aa, the external force, we again separate
out the first term in the summation to have
N N N
T 3,ext = f dl,^1
11
n
/9 = 2
d3Xgd3V0 +
0
2f = 2
df™
dVai
[]
0 =2
d3xg d3v0
(6.2.6)
If we assume that A^- is independentf of vpi, then each term in the sum
again involves an integration of the form
+"
dvai = fw no summation over i (6.2.7)
dl’ai voti = — 00
d/(1)
T 3,ext — Au (6.2.8)
dvu
N
dfw
3,int CLiyi
dVu
P[ d3xp d3Vp (6.2.10)
0 =2
since for any other choice of a we obtain another integration like that
indicated in Eq. (6.2.7), which is zero. In Eq. (6.2.10) we can again
take the derivative out of the integral and integrate over all coordi¬
nates except those of particle 1 and particle 7 to find
In addition, for a system of N identical particles the last term here (in
the limit as N and V —> °°) reduces to
(6.2.14)
For this case, as there is no longer any need to specify any particular
particle, we can also replace xi, vi and x2, v2 by x, v and x', v', so that
the equation for /(1)(x,v,<) becomes
Problem
6.2 Prove (by integrating the Liouville equation over all variables
for the particles from n + 1 to N) that the BBGKY equation for
an arbitrary /(n) is given by
n N
Probably the simplest method one can use for working with the
BBGKY equation is to assume a functional relationship between /(21
and /(1) in the interaction term in the /(1) equation. When this
approach is used, the assumption generally made is that the probability
functions for the two particles are completely independent. For this
case, as we have noted in our discussion of the Boltzmann collision
term in Sec. A.3 of the Appendix, the joint probability that particle 1
is at xi with velocity Vi and particle 2 is at X2 with velocity V2 is just
the product of the two individual probabilities so that
(Note that this is like flipping two coins and asking for the probability
that both turn up heads. If the behavior of each coin is independent
of the other, then this probability is just the product of the two sepa¬
rate probabilities, | X i = }■) If we use this assumption, we imme¬
diately reduce Eq. (6.3.1) to an equation involving just f(1), so that
the truncation of the set of equations is achieved.
One interesting result which follows from this method of trunca¬
tion is that one can reduce the /(1) equation for a plasma to the BV
equation, used in Chap. 4. To see this, let us consider the simple case
of a plasma of N identical electrons in a uniform background of sta¬
tionary ions with number density N/V (to maintain overall charge
neutrality). If we consider just the electrons, then as N and V
approach infinity, we have, from Eqs. (6.2.15) and (6.3.1),
Because of the infinite volume we shall neglect any real external forces.
However, because of the ion distribution, we set
A = — — Eion (6.3.3)
m
where Eion is the electric field produced by the ions, so that (from
Poisson’s equation)
eN
V • Eion = (6.3.4)
V £q
To evaluate the interaction term in Eq. (6.3.2) we note first that for
electrons moving slowly (compared to the speed of light), to a very
good approximation,
e2R
(6.3.5)
Aire^mR3
with R = x — x' the vector pointing from x' to x as shown in Fig. 6.1.
If we now substitute this expression for a into Eq. (6.3.2) and multiply
through by N, we have|
e_df_
dl + v .v Ei I f«y) Ri d3x' d3v' = 0 (6.3.6)
dt ^ 1 dXi m dVi 4 TTti R 3
f Recall that A/(1) is the average value of the usual distribution function /.
However, to reduce the notation we shall here let / d3x d3v be the average number
of particles at x with velocity v.
6.3 THE BV EQUATION WITH SELF-CONSISTENT FIELD 301
Comparing the last term here with the analogous term —{e/m)Ei
df/dVi in the BY equation, we see that they are equal if
e Ri
Et = E*°” - ~ //(<>') d3x' d3v' (6.3.7)
47T€o R3
pc
Eic d3x' d3v' (6.3.8)
47re( / R*
1_ R
V (6.3.9)
R R3
V • Eion + v— f
47T6o J
v2 4
H
d3x' dZv' (6.3.10)
Now we have shown (in Sec. A.l of the Appendix) that for an infinite
volume
n ( x x'
1. 5(x “ x } = (
0
00 = x'
(6.3.12)
f 1 x in the domain of
(6.3.13)
2. /S(x — x') dzx' = <
u /(x)
integration
otherwise
x in the domain of
3. J/(x')5(x - x') d*xr = ■ integration (6.3.14)
_ 0 otherwise
= 0 (6.3.16)
dt ' dXi m di>i
In summary, we see here that the BV equation for an electron gas can
be derived from the BBGKY equation for /(1) when correlation effects
are neglected. Thus, although it includes 110 collision term explicitly,
the BV equation actually includes (in the E-field term) some of the
interparticle effects.
Problem
then for the simple electron gas we considered in the last section we
find [from Eqs. (6.2.15) and (6.3.7)] that
where, again, the prime denotes that the term y = a is omitted in the
sum; while, from the same problem, the equation for/(2)(l,2) is
(6.4.5)
(Note that in the third term here, when a = 1, /3 = 2, and vice versa.)
In the first two terms of Eq. (6.4.5) we now replace/(2) by/(1)(1)/(1)(2)
+ P(l,2). In the first term, we have
V df&
dP‘ V
> vai -— = > vai /(D(/3) dfa)^ + — (6.4.7)
Lt
a=l
& 9^a
dXai
/S
Lt
(=1 ^ ' dxat ^ dxai
Now note that if we add together just the two underlined terms in
Eqs. (6.4.6) and (6.4.7), which are both proportional to /(1)(/3), we
have, from Eq. (6.4.4),
f} 7^ a
2 N
To obtain the full equation for P, we replace the first two terms in
Eq. (6.4.5) by the expressions given in Eqs. (6.4.6) to (6.4.8). In
addition, in the summation on the right side of Eq. (6.4.8) we separate
out the term where y = P, so that the remaining terms run from
y = 3 to 7 = N. With these substitutions Eq. (6.4.5) becomes
a= 1 ' - —
0 ?*«
+ 2 dT-J
7=3
«^l/H>(«,S,T)-/<I>((3)/<!'(a,7)]<i%4xJ -0 (6.4.9)
so that Eq. (6.4.9) now involves only/(1) and P (or /(2)). Even with
this truncation, however, the equation is still too complicated to use.
Hence two additional simplifications are generally made. First, in
the term (with a single underline) involving df(2)/dvai in Eq. (6.4.9) it
is assumed that correlation effects are small, so that
7=3
(6.4.12)
Problem
g—J)a2/a2
/(1)(a) 7r3/2Ea3
and
ele~
Peq(a,P) = -/(1>(a)/(1>(d) R = Ix/s — xa
4-ireoKTR
A De~-RIXD
-/(1)(«)/(1) 08)
"3
with ne = N/V.
f The ratio is actually a function of the distance between the two particles and
becomes much smaller at small separation distances, which is where one would
expect correlation effects to be large. Hence, as we shall note shortly, it is some¬
times important to retain the full expression for /(2) in this term.
6.4 CORRECTIONS TO THE BV EQUATION 307
or
which vanishes when the limits approach 00, since the integrand is an
odd function of Xayi•
Because of these identities, the equations for /(1) and P [Eqs.
(6.4.4) and (6.4.12)] reduce to just
N
and
2 ,
dP , d/««(a)
2
+ a= 1 '
& 9^a
Vai
dXa
— +
dv0
[f(l)(P)aa0i
These equations, while still quite complicated, can be used for some
plasma problems. Probably the most notable work has been the reduc¬
tion of the equations to a Boltzmann-like equation for /(1) which is
called the Lenard-Balescu equation.f
In another interesting paper, TchenJ has used the full expression
for/(2) in the single-underlined term in Eq. (6.4.9), so that the third
term in Eq. (6.4.17) is, in his work, aapi dfw/dvai. Thereafter, by
solving Eq. (6.4.17) for P in terms of/(1) and substituting the result into
Eq. (6.4.16), he obtains (after a number of approximations) the Fokker-
Planck equation discussed previously in Sec. 5.10.
In this derivation, which is too involved to repeat here, each of the
terms in Eq. (6.4.17) plays a distinctive role. For example, in the
third term, where
a/(1)(«) dP(a,P)
/C1)(d) (6.4.18)
dVai dvai
SUMMARY
To summarize we have seen here that it is possible to extend the simple
BV equation obtained by assuming/(2) (a,/3) = /(1)(«)/(1)(/3)- Toobtain
the extension one must take some account of correlation effects. For¬
mally this can be done quite generally, using Eqs. (6.4.4) and (6.4.9)
for/(1) and P but replacing/(3) by Eq. (6.4.10). As a practical matter,
however, we have seen that such an analysis is extremely complicated
algebraically even for the relatively simple case of an infinite homo¬
geneous medium. It is possible, however, that future developments in
this field of research will provide techniques by which analytical results
f The standard references for this work are R. Balescu, Phys. Fluids, 3:52
(1960) and A. Lenard, Ann. Phys., 10:390 (1960). The derivation can also be
found in two recent texts: Montgomery and Tidman, op. cit., and T. Y. Wu,
“Kinetic Equations of Gases and Plasmas,” Addison-Wesley Publishing Company,
Inc., Reading, Mass., 1966.
\ C. M. Tchen, Phys. Rev., 114:394 (1959).
6.5 FORMAL PERTURBATION TECHNIQUE FOR WEAK CORRELATIONS 309
/(3) (1,2,3) = /(1) (1)/(1) (2)/(1) (3) (1 + 012 + 0i;i + 023 + 0123)
=1 f)=a+l 7 = 0+1
In this system of equations, which is called an Ursell-Mayer cluster
expansion, there are three points to note: (1) we have made no approxi-
310 THE BBGKY EQUATIONS FOR A PLASMA
_ P(a,P) (6.5.3)
9a* /(1)(«)/(1)(d)
sr (6.5.7)
/(1) + Var
dX ai
Y f a“^<2)(a’T) d*Xy d3Vy
and
P+j
L,
L~+a.fi^
dXai OVai
-/“‘(ft 4- I a„eir(c,fi)d*x,d>ve
OVai J
a=1 '
P 7^a
+ f J- Jf aayi[fllK«)P(fi,y)
dVai
+f“Ky)P(c<,P)}d3Xyd* V-y] = 0
7 =3
(6.5.8)
Nd = imr\D3 » 1 (6-5,9)
term in Eqs. (6.5.7) and (6.5.8) to make certain that we eliminate any
other terms that are of lower order than iV#-1. For this purpose we
shall apply some simple dimensional analysis to examine the relative
magnitudes of the various terms in these equations.
As is always the case, the choice of a “characteristic” dimensional
quantity is rather arbitrary. However, for the electron gas to which
we generally apply these equations, it is at least plausible to use
the inverse of the electron plasma frequency, as a characteristic time
and Ad, the Debye length, as a characteristic distance over which /(1)
or P might change. (Recall that we have, in fact, already used Ad as
a characteristic length when we assumed gals was of order No-1.) To
obtain a characteristic velocity, we recall that, as shown in Eq. (2.4.20),
(6.5.10)
which is of the order of the average thermal velocity for the electrons.
We emphasize again that for some problems (involving beams, for
example, where the beam velocity greatly exceeds vth, or involving
structures with dimensions smaller than Ad) we might well use a different
set of characteristic values. However, the choice we are using here is
the most reasonable one for large relatively homogeneous stationary
plasmas.
With these characteristic values for time, length, and velocity, we
can now estimate the order of each term in Eqs. (6.5.7) and (6.5.8).
In the /(1) equation, for the first two terms we have
(6.5.11)
so that these terms are of comparable importance. Next, for the inter¬
particle force terms, note that for a coulomb force (assuming the dis¬
tance between particles is of order Ad) we have
<9/(2) g2«2)
aayiW~- ~ -7 2- (6.5.12)
dVai 4:irme0\D2Vih
To put this into a form which can be more easily compared with
the other terms, we use Eq. (6.5.10) to eliminate vlh and recall that
we2 = ne2/me0, so that Eq. (6.5.12) reduces to
(6.5.13)
6.5 FORMAL PERTURBATION TECHNIQUE FOR WEAK CORRELATIONS 313
dP Vth
P '—' coeP P CO ,p (6.5.14)
’ dXai
so that these two terms are of the same order, but both are of order
ND~l compared to such terms as coe/(1)(a)/(1)(/3). Of the remaining
terms in Eq. (6.5.8) all involve the operator
d We (6.5.15)
dafU
dV0 3 ND
^+1 a—1 1
P
{■ dXai
+ /(1)(£) d~^~- (a«* - <a^»
OVai
N 1
= 0 (6.5.16)
+ I aT / a“^(1)(a)p^’T) +f(1)(i)PM]d3xyd3v^
7=3
where
(6.4.17), the equation used by Lenard and Balescu (as well as many
other workers). In summary, what we have seen in this section is that
an ordering of terms with respect to Ncan be used to obtain a pair
of equations for/(1) and P, rather than merely relying upon the ad hoc
procedures of the last sections. Although modifications of this sort
provide a somewhat more satisfactory basis for solving the BBGKY
hierarchy of equations, it is clear that they are still far from being
rigorous or complete. Hence these kinetic equations must be treated
as tentative until they have either been tested against more experi¬
mental results or put onto a more sound theoretical base.
appendix
MATHEMATICAL
APPENDIX
appendix
( 0 for X ^ X'
5
1. 8(x — x') = (A.1.1)
j 00 for x = x'
2. [X+€
Jx-t
S(x - x') dx' = 1 (« > 0) (A.1.2)
The part of the integrand proportional to the cosine is even, while the
term proportional to the sine is odd; hence we have
. 1 /* 00
8(x — x') = lim - / e~ak cos k(x — x') dk
a—*0 TT
(A.1.6)
a-*0 7T «2 + (x — x')2
A.l DELTA FUNCTION AND FOURIER TRANSFORM THEORY 317
a fx+t dx' __ a fe dz _ 2 1
(A.l.7)
TV Jx-< a2 + (x - x'Y ~ TV J-e a2 + z2 “ tv 1111 a
where F(x') is given in Eq. (A.l .9) and e is a very small positive number
which is larger than a but which also approaches zero. F rom Eq.
(A.l.9) the first and third integrals vanish, while in the second integral
we again let z = x' — x, to have
af(z + x) (A.l.11)
lim dz
0 7r(a2 + z2)
a dz (A.l.12)
jj = f(z + x) and dV
tv (a2 + z2)
318 MATHEMATICAL APPENDIX
so that
J* Fix') di
(A. 1.15)
(A. 1.16)
Although this procedure is not strictly valid, it does simplify the nota¬
tion a great deal, and for all problems of physical interest it leads to the
same results that one would obtain using the limit shown in Eq. (A. 1.4)
or (A.1.9).
As an example of the usefulness of this <5 function representation,
we can easily prove the following basic theorem in Fourier transform
theory. Given
(A.1.17)
(A.1.18)
A.l DELTA FUNCTION AND FOURIER TRANSFORM THEORY 319
To prove Eq. (A.l. 18) we need only compute the integral on the right.
Using Eq. (A. 1.17) for/(re), we have
Substituting this result into Eq. (A. 1.19), which means, in essence, that
we have interchanged the order of integration, and applying condition 3
[in Eq. (A.1.3)], we have
Problem
A.l Prove the three-dimensional Fourier transform theorem stated in
Eqs. (A.1.22) and (A. 1.23).
One way to prove Eq. (A. 1.24) would be to verify directly that
the function on the right satisfies all the delta-function properties.
However it is just as easy to instead show that the differential equation
can be solved to find 4> = (4irR)-1. To solve Eq. (A. 1.26), we use
Fourier transform theory to first find $(k).
If we replace 4>(x) by its Fourier transform,
Conversely, on the right-hand side of Eq. (A. 1.26) we can replace each
of the delta functions by an expression like Eq. (A. 1.16), to have
Equating Eqs. (A.1.28) and (A.1.29), we find that the Fourier trans¬
form of 4>(x) is given by
e-ik-x'
$(k) = (A.1.30)
8tt3/c2
J r gik-R
4>(x) = f 4>(k)e,k'x d3k
8V* J ~Wd3k (A.1.31)
A.2 EVALUATION OF SCATTERING ANGLE x 321
(A.l.33)
nm = - 1| (a.2.2)
F
R = *i - * = - (A.2.3)
with n = mmi/(m + rax) the reduced mass. Note that this is the equa¬
tion of motion for a particle of mass g, whose position and velocity are
that of particle mx relative to m, and which is acted on by the force
F = — d$/dR. The geometry of this problem is shown in Fig. A.2,
where we choose the origin of the coordinate system to be the point
where R = 0, and where we set a„ perpendicular to the plane of the
orbit. Note that the impact parameter b would be the distance of
closest approach Rm if the two particles did not interact. In addition,
R (which always points in the direction of the path of the orbit) is the
relative velocity of the two particles. Hence, the scattering angle x
f Note that we neglect here any effect on the particles of external forces.
This is due to the fact that the interparticle forces generally act over just a very
short range but, in this close range, the forces are very large compared to typical
external forces. In addition, as a practical matter, the solution to the collision
problem where there are large external forces is so complicated that one would
probably not bother treating such problems via the Boltzmann equation but would
instead use the BBGKY equations of Chap. 6.
A.2 EVALUATION OF SCATTERING ANGLE x 323
between g and g' is the angle between the two asymptotes of the path
and can be easily found if we know R(6).
The path is most easily found by using the conservation laws for energy
and angular momentum.f First, setting the kinetic and potential ener¬
gies at any point equal to the initial kinetic energy (since the initial
potential energy is zero), we have
L = m(R X R + R X R)
which is zero, since, from Eq. (A.2.3), R and R are parallel, so that their cross
product (like R X R) vanishes. Hence, for this case L is constant, and the path
lies in a single plane.
324 MATHEMATICAL APPENDIX
x = R sin 9 z = R cos 9
(A.2.5)
x = R sin 9 + RO cos 9 z = R cos 6 — Rd sin 9
Hence, we find
R • R = x2 + i2 = R2 + R2 A2 (A.2.6)
m = b9 (A.2.10)
so that
bdR
dd (A.2.12)
± R2X
A.2 EVALUATION OF SCATTERING ANGLE 325
A.=b (i - <A-2-i3>
Evaluation of x
From Eq. (A.2.12) we can compute the orbit and the scattering angle x-
First, however, we note that the orbit is symmetric about the angle dm,
since if we integrate Eq. (A.2.12) from 9m to some other angle 9, we have
rR b dR' (A.2.14)
6 ~ 9m = ± ]Rm Rnxm
where the plus sign is used for 9 > 9m and the minus sign for 9 < 9m.
Note that when R —> °o , —> 0 while 9+ > 2(?m. However, from Fig.
A.2, the latter asymptote is also tv — x) hence we have
(A.2.15)
X = tt ~ 2 9m
or
of-__ (A.2.16)
x _ T “ Jr, 7f2[l - (b/R)2 - 2(4>/w72)]1/2
0 R > a (A.2.17)
HR) = 00 R < a
f Note that the monotonic increase of 0 with time follows from Eq. (A.2.10),
since R, b, and g are all positive.
326 MATHEMATICAL APPENDIX
(b/Rm)
= *-2 /„* (Id 7T 2 sin —i (A.2.18)
Rr
If b exceeds a, it is clear from Fig. A.2 that the particles never collide,
and Rm = b, whereas, for b < a there is a collision, so that Rm = a.
Applying this to Eq. (A.2.18), we have
0 ■ _tb
7T — 2 sm 1 - b < a
X a (A.2.19)
o b > a
Problem
F = ^
47re0E2
Evaluation of AN
In Sec. 4.1 we derived the equation for the rate of change of the velocity
distribution function /
dfa
dt + Vifv
oXi + a^
dVi (A.3.1)
where the term on the right is the rate of change of fa due to collisions.
More specifically if we let AN denote the net change (due to collisions
during time At) of the number of particles of type «, /„ dzx dh>, in the
volume element d3x at x with velocities in the velocity-space volume
element d3v at v, then as in Eq. (4.1.15)
AN = d3v A<
(A.3.2)
A.3 DERIVATION OF BOLTZMANN COLLISION TERM 327
by the product
If we use this identity in the expression for AJV+, we have, from Eqs
(A.3.2), (A.3.4), and (A.3.6),
MA = AN+ - AN-
\ dt Jc d3x d3v At
• X (A.3.9)
sin
2 (1 + no2)1'2
(A.4.1)
b>a
with o- the sum of the radii of the colliding spheres. Hence the
integrand in Eq. (A.4.1) is zero for b > cr, and we have for hard-sphere
collisions,
To complete the calculation, note [from Eq. (A.4.2)] that over this
range of integration
so that
db = -fr sinixdx
l odd
lira2
(A.4.6)
l even
J+ 1
Thus, for odd values of l this cross section is just the geometric
cross section ira2, while for even values of l this cross section is slightly
smaller.
K (A.4.7)
(p — 1 )RP~1
332 MATHEMATICAL APPENDIX
Using this potential in the formula for the scattering angle x [Eq.
(A.2.16)] and introducing two nondimensional variables /3 and Vo,
we find
2 r R™ dV-1 -1/2
d/3
X = 7T 1 - /32 (A.4.8)
JO p — 1 \v0/
where (3 = b/R [so that d/3 = — (b/R2) dA], and w0 is chosen so that
(A.4.9)
/jcyi(p-i)
SaP^ = 2ir Mv) (A.4.10)
U2/
qaqfs (A.4.12)
=
4ireoR
O/.% 2 yyt
4>d = <t>c exp ( — Icd'R) kD>2 = (A.4.13)
provided the impact parameter is less than Ad, while if b exceeds Ad,
we shall assume that x = 0. In other words we neglect the screening
for b < \D and assume perfect screening for 6 > \D- With this
assumption
Vo2 ~ 1 b < Ad
2 sin2 ^x = | v02 + 1 (A.4.15)
cos x =
b > Ad
Vo2 - IV
(A.4.16)
Sa^ = 2tr b db
Ho2 + 1/
f Note that this is the total potential energy for the two charges in contrast
with the coulomb potential per unit charge used in Sec. 4.8, but also denoted by
334 MATHEMATICAL APPENDIX
4nreoMfi^Y
(A.4.17)
-it r>C2\d
Saf)v = 1 - (A.4.18)
= C2 Jo
(A.4.19)
= 2tvC-H In (z + 1) + t-i
(i - m - 2) C*\d
0 + 1 3(0 + l)2 0
(A.4.20)
which does not diverge as a result of our having cut off the integration
at 6 = \D. (Otherwise, the upper limit here would be infinity and the
log term would diverge.)
In view of the approximate nature of this calculation (due to the
uncertainty in the actual interparticle potential), it is customary to
simplify this result still further by replacing the quantity in the upper
limit (C\D)2 (which is a function of g2) by its average value. If we
assume that the a and gases have no drift velocity and a common
temperature, then we find by direct calculation that
Applying this to the expression for CXD, we find that for singly charged
particles
12™ Ten
A = (C\D) = / = 127rneXn3 = 9 ND (A.4.22)
e2
(i - m - 2)
3
(A.4.23)
This is the cross section commonly used for problems involving coulomb
collisions. Although it is possible to make calculations which are
much more sophisticated than this, the final results are always virtually
identical to this, except possibly for differences in the slowly varying
function In A.
Problem
A.3 Find the average value of g2 and calculate A for the case when
species a and /3 have arbitrary distribution functions and where
Ta 9^ Tp but ua = up = 0.
Lowest-order results
As noted in Chap. 5, the original technique developed for solving the
Boltzmann equation and calculating the transport coefficients for a
gas is due to Chapman and Enskog. Their method is an expansion
procedure based upon the limit when the gas is close to equilibrium and
the collision frequency is very high. In order to discuss the method
as simply as possible, we shall restrict our attention to the Boltzmann
equation for a simple electron gas and use the Krook collision model,
so that we have
(A.5.1)
ne-cVa2
2 kT
c = v — u (A.5.2)
7r3/2a3 m
so that (as in Secs. 4.3 and 5.7) the collision integrals for mass, momen¬
tum, and energy transfer are all zero.
The basic expansion procedure used in the Chapman-Enskog
technique is to assume that the self-collision frequency v is very large
and that / can be expanded in a series of the form
/ = Ao + V~1A\ + r~2A2 + • ■ •
In this expansion all the A/s are assumed to be of the same order, so
that
Dn
+ nV • u = 0 (A.5.6)
Dt
Duk 1 dpik
~Dt - {ak) ~ p (A.5.7)
and
If we substitute these expressions for qk(-0) and pt-/0) into the conserva¬
tions, we obtain the Euler equations of hydrodynamics
Dn
— nV • u (A.5.11)
JTt
Duk . , 1 dp
(A.5.12)
Dt ~ p dxk
and
D — 5/3
0 (A.5.13)
DtVU
There are several alternate forms for the last of these equations (the
energy equation). For example, if we replace p by ukT and use Eq.
(A.5.11) for Dn/Dt, we find
DT
-fTV•u (A.5.14)
Dt
t Note that Eq. (A.5.10) has been calculated earlier [see Eq. (5.7.4)], while
Eq. (A.5.9) follows from the fact that the integrand is odd in ck.
338 MATHEMATICAL APPENDIX
Calculation of /(1)
To calculate the next term in/ we consider the terms in the Boltzmann
equation (A.5.1) that are independent of v
dAo . i dAo dA o
+ U; — A\ (A.5.15)
+ Vi ~dxl dVi
S)n _ Dn dn
iDt ~ Dt + Ci dXi
£>Cj /Dm dui _
SD t ~ \Dt + Cj dxj " (A.5.18)
2D n dn
■nV • u + c,
dxi
2D Ci 1 dp
- (e, X B); (A.5.19)
p dXi m
2D<3
2D t
1 2DAc _ _/oA
(A.5.20)
r 2D t v
with
A = | W - I) ^ - icHti) || (A.5.21)
dT 5 up
5*C1) = — Xo (A.5.23)
dxk 2 mv
dUk
Pki(l) = Vo (%hiV ■ u (A.5.24)
dxi
Duk
Dt
£m ia + (u X B)„] p dXk
+ mp V2uk + l
6 dXk
(V • u)
(A.5.25)
p5/3 R_
r -FT,
Fw PP-5/3 (A.5.26)
Lowest-order analysis
and
To = 0 Ti — -Mn„Sen^(v) (A.6.6)
Next-order treatment of T0
The evaluation of T0 to next order is a difficult calculation which is
very similar to the calculation of the Fokker-Planck coefficients in
Sec. 5.10. As in that section, the analysis can be shortened somewhat
by introducing a new form for T0. Suppose we multiply both sides
of Eq. (A.6.4) by some arbitrary function Q(v) and integrate both
sides over all values of v. Then, upon interchanging primed and
unprimed quantities in the first term on the right and recalling that
g' = g and d3v' d3V' = d3v d3V, we have
We now note from Eq. (A.6.4) that although g, v', and 1 ' all
depend on the angle between v and V, there is no term in To that
depends on the orientation of v relative to the laboratory coordinate
system. Hence, after we integrate over all values of V, both 7’0 and
the integrand on the right side of Eq. (A.6.7) must, by symmetry,
depend only on the magnitude of v. For this reason, it we write d3v in
spherical coordinates and integrate both sides of Eq. (A.6.7) over the
342 MATHEMATICAL APPENDIX
two angles, we have (after canceling the factor 4-ir on each side)
v' = v + Ay (A.6.9)
so that we have
Jo QT0v2 dv d2Q p / u
(A.6.11)
^2 G^V) dv
where
Since this equality must hold for any function Q(v), it follows that
1/ rfGj d2GA
T
J o
v2 \ dv dv2 J (A.6.14)
m
m0 = m + M (A.6.16)
v' = v + ^'-*>
2w?
y2 = v2 + —
m0
V-(g'-g)+^ 02(1 - cos x) (A.6.18)
Trio
Upon using this expression for V'2 in Eq. (A.6.15), we have with
no approximation,
m0 1 + m/M 1 M
(A.6.20)
m ( 1 V^ m
m0 M \1 + m/M) ~~ M
2m
./2
i +4v (g - g')
M v2
g2(l - cos x) (A.6.21)
where the first underlined term here is of order (m/M)112, and the
second is of order m/M. Therefore, if we take the square root of
both sides of Eq. (A.6.21), using the formula
we find
v' — v + Av (A.6.22)
Av
— v~ {v- (g - g') -j„'( 1 - cos x) - [V' (S2[J g,)p
v (A.6.23)
/ \ 3/2
—J = y-4[y . (g _ g')]2 + terms of order (A.6.24)
Now that expressions (to order m/M) have been obtained for Av and
(Av)2, we can proceed with the calculation of G^v) and G2(v) [as given
in Eq. (A.6.12)]. To perform the integrations over b and e we recall
that in the text we show that [see Eqs. (5.4.25), (5.4.30), and (5.10.17)]
with
^ij 'V/i(lj
Uij (A.6.29)
g gz
Therefore, when we use Eqs. (A.6.23) and (A.6.25) for Av and (Ar)2
in the expressions for Gi(v) and Gi{v) [Eq. (A.6.12)] and integrate over
b and e, we find
and
gi - —’Vi + Vi
(A.6.32)
(1 +e)i/2 = 1 +*e-
we have
vkV k m\
(A.6.33)
« = t'{1- — + eM)
(A.6.34)
Since the term Gt(v) contains the factor VlVJ- (which is of order
mv2/M), the expansion to order m/M in this term involves the single
346 MATHEMATICAL APPENDIX
M
8ijnnA 2
2 (A.6.35)
and letting
X (TinSen) (A.6.36)
fov3A2
G2(v) (A.6.37)
A2 mv*
Gt(v) = fo
2 MX
, /A2 d, v3 m z>4\
h \Y dv \ ~ M X ) (A.6.40)
Final result
From these expressions for G^v) and G2(v) we can now calculate
T0(v) = f(f'0F' - f0F)gb db de d3V. From Eqs. (A.6.14) and (A.6.37)
we have
The underlined term here exactly cancels the first term in Gi(v) [note
Eq. (A.6.40)], so that we find
m A2 d_ /v3 df0\
(A.6.42)
Mv2 dv X 2v2dv\\ dv)
Physical constants
Conversion factors
Plasma parameters
Adiabatic invariants: p = A
Ah oi
= 2 ir
\vx\L
B<
Free-electron-diffusion coefficient: De
ITleV e
x(Te + Td
Ambipolar diffusion coefficient: DA
meven + mtvi
CTe + Ti)
miVi
A.7 IMPORTANT FORMULAS AND DEFINITIONS 349
Heat-flow vector, definition, 163, 164 Kinetic-energy density, 110, 111, 249
Kirkwood, J. G., 292
in Euler approximation, 242, 243, 337
Kronecker delta, 168
in magnetic field, 269, 270
Krook collision model, applied to Chap-
in Navier-Stokes approximation, 339
man-Enskog technique, 335-340
340
applied to Landau damping, 181, 189,
related to 13-moment coefficients, 260
197
263
description of, 173-177, 224
Holes, 1, 8, 92, 133
in stability problems, 200
Hydrodynamic derivative, 105, 155, 169
with 13-moment equations, 260, 261,
Hydrodynamic viscosity (see Viscosity) 265, 268, 273.
INDEX 359
4 » 'gf
y 4
iiOV ££
0 1163 0148533 4
TRENT UNIVERSITY
QC718 .T23
Tanenbaum, B Samuel
Plasma physics
113306
■v