You are on page 1of 335

Fundamentals of

Plasma Physics
James D. Callen
University of Wisconsin, Madison
June 3, 2003

PREFACE
Plasma physics is a relatively new branch of physics that became a mature
science over the last half of the 20th century. It builds on the fundamental areas
of classical physics: mechanics, electrodynamics, statistical mechanics, kinetic
theory of gases, and fluid mechanics. The distinguishing feature of the plasma
medium is that its properties are determined by the nature of the interactions
between the charged particles in it collective rather than binary and weak
compared to their thermal motions.
The collective but weak interactions in a plasma embody many physical
processes over a wide range of length and time scales: predominantly deterministic particle motion which however may be diffusive on long time scales,
internal generation of microscopically irregular but macroscopically smooth electromagnetic fields, both adiabatic and inertial (or fluidlike) plasma responses,
dielectric medium type electrical properties, and various flow regimes (laminar,
transitional, shock and turbulent). These processes lead to a wide variety of
interesting collective phenomena, e.g., dielectric shielding of charges, waves in
the medium, transfer of energy from waves to particles (via Landau damping, a
collisionless wave-particle resonance effect), transfer of energy from a distribution of particles into waves (instabilities), and turbulence in the six-dimensional
(three real plus three velocity space coordinates) phase space.
Increased understanding of plasma physics has both been stimulated by, and
paced, the development of its many important applications, e.g., magnetic and
inertial approaches to fusion, space and astrophysical plasmas, plasma processing of materials, and coherent radiation generation (typically via acceleration of
beams of electrons or ions). Thus, plasma physics has developed in large part
as a branch of applied or engineering physics science with a purpose.
The primary objective of this book is to present and develop the fundamentals and principal applications of plasma physics, with an emphasis on what
is usually called high-temperature plasma physics where the plasma is nearly
fully ionized with nearly negligible effects of neutral particles on the plasma
behavior. The level is meant to be suitable for senior undergraduate students
through advanced graduate students and active researchers. Pedagogically, it
begins from an elemental or microscopic description, then uses this to develop
macroscopic models of plasmas, and finally uses these models to discuss practical applications. A variety of applications of plasma physics are discussed
throughout the text; many others are covered in the problems at the end of
each chapter. In concert with the modern trend in the physical sciences, SI
(Syst`eme International dUnites) or mks units are used throughout.
This book has evolved primarily from lecture notes developed while teaching various plasma physics courses at the University of Wisconsin-Madison over
more than two decades (19792003) and in part from teaching three years at
Massachusetts Institute of Technology (19691972). My own research and teaching has been predominantly in magnetic fusion research, which has been the
dominant driving force behind the development of the science of plasma physics
over this period. However, because plasma physics has grown into a mature

ii

PREFACE

science whose principles are broadly applicable, I attempt to develop the fundamental concepts in an application-independent manner. In addition, many
different types of applications of plasma physics are discussed throughout the
book.
The science of plasma physics draws heavily on many areas of classical
physics and applied mathematics. Typically, not all of these subjects are well
known to the wide variety of students (from physics, engineering physics, electrical engineering, nuclear engineering and other undergraduate curricula) who
begin studies of plasma physics. Also, most of the needed background material
is not readily available in concise, accessible forms. Thus, a number of Appendices have been written to provide relevant summaries; they give important
supplementary information that is an integral part of this textbook. Finally,
Appendix Z (to be placed on pages inside book covers) provides sets of basic formulas that are useful throughout the book vector relations, vector
differentiation operators, physical constants, and key plasma formulas.
This book is designed for teaching plasma physics at a variety of levels.
(It may also serve as a useful reference book for active researchers in plasma
physics.) For example, it could be used as the basis for a two (or more) semester
graduate-level course on plasma physics, at the rate of approximately one chapter section per one hour lecture. However, it could also be used for teaching
a fast-paced, one-semester introductory course on plasma physics by covering
only the sections at the first of most of the chapters. Intermediate-level subjects
that could be omitted without compromising understanding of later sections are
indicated by an asterisk (*) at the end of the respective section titles. Advanced
material, which is relevant mostly for research purposes, is similarly indicated
by a plus sign (+). Bibliographies at the end of each of the chapters and appendices provide information on other textbooks and research literature that should
be consulted for further details or supplementary course material. Individual
chapters of this book will be made available (in draft form) via my public web
page (http://www.cae.wisc.edu/callen) as soon as they are available.
The large number of problems at the end of each chapter are graduated in
level of difficulty commensurate with the various levels and styles of courses that
might be taught from the book. Specifically, the levels of the problems are classified according to their nature and consequent degree of difficulty: evaluational
(/), application development (//) and conceptual development (///). Also, the
level of material involved in solving the problem is indicated: basic (no mark),
intermediate (*), or advanced (+).
(Detailed acknowledgements of help by others and assistance in the preparation of this manuscript will be written later.)

DRAFT 10:38
June 3, 2003

c
"J.D
Callen,

PREFACE
Fundamentals of Plasma Physics

iii

Introduction
Plasma is often called the fourth state of matter. The various states of matter
occur as a substance is heated to temperatures above the binding energies for
particular states of matter and thereby undergoes phase transitions. As an
example, consider the states of H2 0 and its molecular, atomic and elementary
fig:1
particle constituents at various temperatures, as indicated in Fig. 1. Below
273 K ( 0.0235 eV1 ) it is in a crystalline form known as ice a solid, the
1 Temperatures (and particle energies) in plasma physics are usually quoted in electron
volts, abbreviated eV. The conversion factor from Kelvin to eV is Boltzmanns constant kB
divided by the electron charge: kB /e ! eV/11,604.4 K.

Temperature
K

States
of H2O

States of
Matter

Types!of!Plasmas

eV

1.16!!1013

1!GeV
nuclear
plasma

1.16!!1010

1!MeV

relativistic plasmas
plasma

1.16!!107

1.16!!10

1!keV

1!eV

fully
ionized
plasma

partially
ionized!gas
steam

(H2O!molecules)

water
ice

373
273
11.6

neutral
gas
liquid
solid

fusion reactors
fusion experiments
interior of sun
solar corona
earths magnetosphere
solar wind
gas lasers
fluorescent!lamps,
gaseous electronics
earths ionosphere
interstellar space

.001!eV

Figure 1: Schematic of states of H2 O as it is heated. Also shown are the


corresponding states of matter and some of the types of plasmas that can occur
in the various temperature ranges indicated.

DRAFT 10:38
June 3, 2003

c
"J.D
Callen,

INTRODUCTION
Fundamentals of Plasma Physics

fig:1

iv

INTRODUCTION

first state of matter, which is a strongly coupled medium (binding energy large
compared to thermal energy). At temperatures between 273 K and 373 K the
crystalline bonds are broken, but large scale molecular structures exist and H2 0
is called water a liquid, the second state of matter, which is also a strongly
coupled medium. At temperatures above 373 K ( 0.032 eV) the long-scale
molecular structure bonds are broken and the independent H2 0 molecules form a
gas, which is commonly known as steam. Upon further heating to a temperature
of the order of the molecular binding energy ( 0.3 eV), the molecules dissociate
into independent hydrogen and oxygen atoms. While this is no longer steam,
it is still a gas in which the elemental constituents (H2 and 02 ) are electrically
neutral. This third state of matter is a neutral gas, which is a weakly coupled
medium on average, interactions between particles are weak, compared to
their thermal motions.
We finally reach the plasma or fourth state of matter when we heat the gas
to the point where a significant fraction of the atoms are dissociated (atomic
bonds broken) into negatively charged electrons and positively charged ions to
form an ionized gas. The fraction of the atoms that are dissociated is called
the degree of ionization. The binding energy of the most weakly bound electron
in atoms of all types is typically of the order of the 13.6 eV binding energy of
an electron in the hydrogen atom. As discussed in Section 7 of Appendix A
(Section A.7), when the temperature increases to a significant fraction ( 0.02
1) of the electron binding energy, collisions between the atoms in their thermal
motion cause a nonnegligible fraction of the atoms to become ionized. Electron
temperatures in the few eV range typically produce a partially ionized gas.
An ionized gas is in the plasma state if the charged particle interactions are
predominantly collective rather than just binary. (Binary interaction collisions
are one-at-a-time interactions with other individual charged particles or neutral
atoms and are the dominant ones in neutral gases.) In a plasma the interactions are collective because many charged particles interact simultaneously,
but weakly, through their long-range electromagnetic fields, and in particular
their Coulomb electric fields. Thus, a plasma is a collective but weakly coupled
medium in which interaction energies are much smaller than thermal energies.
At temperatures above a few eV the ionization becomes essentially complete.
At this point, it is an almost completely ionized gas and it is nearly always in
the plasma state; hence it is then usually called a fully ionized plasma. Further
heating of a collection of such particles would successively break nuclear bonds
( MeV) and quark bonds ( GeV). These result in nuclear and quark-gluon
plasmas, respectively. However, such states are beyond the scope of normal
plasma physics and will not be treated in this book.
The word plasma, which comes from the Greek , means something
molded. It was introduced by Tonks and Langmuir in 1929 to describe the
behavior of the ionized gas in an electrical discharge tube, which they found
could be manipulated by a magnetic field. While most plasmas can indeed be
manipulated by magnetic fields to some degree, their collective behavior often
resembles that of an electrically charged, shapeless, structureless fluid that oozes
about mostly of its own accord, as one might imagine an electrically active
DRAFT 10:38
June 3, 2003

c
"J.D
Callen,

INTRODUCTION
Fundamentals of Plasma Physics

INTRODUCTION

lump of jelly would. Thus, the name plasma is only partially appropriate it
expresses a hope but perhaps not always the reality.
fig:1
Some common types of plasmas are indicated on the right side of Fig. 1. Partially ionized plasmas include various types of gas discharges (fluorescent lamps,
gas lasers, arc discharges, plasmas for materials processing) and the earths ionosphere. The earths magnetosphere and the solar corona are prominent space
physics examples of nearly fully ionized plasmas. Since most of the vastness of
interstellar space is in the plasma state, it is often said that 99% of the universe
is governed by plasma physics. (However, since the interiors of stars are also in
the plasma state, the actual fraction of particles in the universe that are in the
plasma state is much closer to unity.)
The most prominent examples of high temperature, essentially fully ionized
plasmas are those in the solar wind and in fusion experiments. The latter
experiments seek to confine plasmas either with magnetic fields or inertially
at temperatures of about 10 keV or greater together with a product of the
plasma density and the plasma confinement time of more than 1020 m3 s. The
objective of creating such plasmas is to develop an environmentally attractive
new energy source based on the exothermic fusion of light ions. For example,
the fusion of deuterium and tritium (isotopes of hydogen) nuclei produces 14.06
MeV of energy, which is much larger than the 4.65 keV of collision energy
that is required to overcome the Coulomb potential barrier. In addition to
these thermal plasma examples, many types of modern devices for generating
coherent radiation are governed by the collective interactions of plasma physics:
free electron lasers, ion beams, relativistic electron beams, and gyrotrons.
This book concentrates on the physics of fully ionized, nonrelativistic plasmas composed of electrons and ions, which usually means temperatures and
particle energies ranging from about 10 eV to 100 keV. The physics of partially
ionized plasmas, which combines plasma and atomic physics, and chemistry, is
covered only partially through a few examples and problems. Quantum mechanical effects are mostly neglected because, while there are various types of
quantum mechanical plasmas, for the plasmas of interest here the most relevant
interaction distances are usually much longer than the de Broglie wavelength.
The fundamental processes in a plasma are governed primarily by classical
physics. The motion and interactions of charged particles are described by the
usual equations of classical mechanics and electrodynamics see Appendix A.
While relativistic effects in mechanics are important for radiative processes and
in very hot, relativistic plasmas where the electron temperature becomes a
significant fraction of the electron rest mass energy (511 keV), they can mostly
be neglected for the plasmas of interest here.
The distribution of the charged particles in the relevant six-dimensional
phase space (three spatial and three velocity space coordinates) is governed
by a plasma kinetic equation that takes account of the motion of charged particles in the extant electromagnetic fields, and of the Coulomb collisions between
the charged particles in the plasma. While the velocity distribution of charged
particles in a plasma is often close to the collisional equilibrium Maxwellian distribution, ordinary statistical mechanics is not usually applicable to plasmas
DRAFT 10:38
June 3, 2003

c
"J.D
Callen,

INTRODUCTION
Fundamentals of Plasma Physics

vi

INTRODUCTION

because collisional relaxation processes in plasmas are quite slow (compared to


various physical processes in plasmas), and because plasmas are often in unstable and hence strongly nonequilibrium states. In unstable plasmas small perturbations grow exponentially in time by transferring energy from the charged
particle distribution into collective motions of the plasma. Non-equilibrium statistical mechanics descriptions have been developed for some particular plasma
situations; however, it has not been possible to give a general description of
plasmas using this approach.
When the velocity distribution is close to a Maxwellian, it is often sufficient
to use fluid moment descriptions (e.g., plasma density, momentum, and energy
equations). Then, the description of plasmas becomes analogous to descriptions
of ordinary neutral fluids. However, the effects of electromagnetic fields on
the charged particles and the separate (and often different) behavior of the
electron and ion components in a plasma make these fluid moment descriptions
much more complicated. Nonetheless, plasmas exhibit a rich variety of the
types of phenomena usually associated with neutral fluids wave propagation,
instabilities, turbulence, and turbulent transport.
This book is organized broadly as follows. Part I develops descriptions of
the fundamental processes in plasmas collective phenomena, Coulomb collisions, structure of magnetic fields, charged particle motion, and the various
models [kinetic, two-fluid, and magnetohydrodynamics (MHD)] that are used
to describe plasmas. Then, Part II discusses the various types of waves that occur in stable plasmas. Plasma kinetic theory and its applications are discussed
in Part III. The plasma transport processes induced by Coulomb collisions in a
stable plasma and their effects on plasma confinement are discussed in Part IV.
The equilibrium and stability properties of a plasma are developed in Part V.
Finally, Part VI provides an introduction to nonlinear plasma theory, and to
plasma turbulence and the anomalous transport it induces.

DRAFT 10:38
June 3, 2003

c
"J.D
Callen,

INTRODUCTION
Fundamentals of Plasma Physics

Part I

Fundamental Processes in
Plasmas

2
A general definition of a plasma is: plasma is an ionized gas or other medium
in which charged particle interactions are predominantly collective. By an ionized gas we mean that there are significant numbers of free (unbound)
electrons and electrically charged ions in addition to the neutral atoms and
molecules normally present in a gas. In a neutral gas the particle interactions
are dominated by isolated, distinct two-particle (binary) collisions. In contrast,
in a plasma the charged particles interact simultaneously and hence collectively
with many other nearby charged particles in the plasma. However, the typical
particle interaction energies are small compared to the thermal energies of the
particles. Thus, a plasmas behavior is determined by the collective but weak
interactions between large numbers of nearby charged particles in it.
Charged particles interact collectively in most plasmas through their electromagnetic fields. In collective interactions many charged particles interact simultaneously because the Coulomb electric field force induced by each charged
particle is a long range force that decreases as only the reciprocal of the square
of the distance from the charged particle. Thus, a test charged particle experiences the sum of the electric field forces from many nearby charged particles.
The interaction is collective because the nearby particles also experience and
respond to the electric field forces from all the other nearby charged particles,
as well as that of the test particle. Hence, a plasma is a highly polarizable
medium. These collective rather than binary charged particle interactions in a
plasma lead to a wide variety of interesting phenomena collective (Debye)
shielding of individual charges, oscillations at the plasma frequency, dielectric
medium responses to perturbations, and wave propagation. Chapter 1 develops
descriptions of these fundamental collective phenomena and their consequences.
Collisions of charged particles in plasmas are quite different from normal
neutral particle collisions. Neutral particles move independently along straightline trajectories between distinct collision events, which are typically strong,
inelastic events that cause the neutral to be scattered in an approximately random direction. In contrast, a charged test particle moving through a plasma
simultaneously experiences (and is deflected by) the weak Coulomb electric field
forces around all the nearby charged particles as it passes by each of them.
Since the electric fields around the individual charged particles are quite weak
and Coulomb collisions are elastic (energy-conserving), they individually lead to
typically only very small deflections in the direction of motion of the initial test
particle. Thus, the trajectory of a charged test particle is influenced by many
simultaneous, small angle deflections in its direction of motion, with occasional
larger deflections when it passes close to another charged particle. Because
charged particles in an ionized gas are usually essentially randomly distributed
in space, the deflections produced by Coulomb collisions are random and lead
to a diffusive or random walk (Brownian motion) process in the direction of motion (or velocity vector) of a charged particle. The properties of the cumulative
effects of many Coulomb collisions on a single charged test particle, including
the effective collision frequency for 90 deflection of its velocity vector, and the
net collisional effects on a near Maxwellian distribution of such particles are
developed in Chapter 2.

3
In some of the most important applications of plasma physics a quasistationary magnetic field permeates the plasma e.g., in magnetic confinement
devices for fusion, the solar corona, and in the earths magnetosphere. These
magnetic fields can have quite complicated behavior (e.g., curvature, shear) and
structures (e.g., magnetic islands). Since we will want to investigate the properties of plasmas imbedded in such magnetic fields, in Chapter 3 we discuss the
general structure (kinematics) of magnetic fields and the mathematical models
(local and global) used to describe them.
Charged particles in plasmas move along trajectories governed by a combination of inertia (m dv/dt = 0 = x = x0 + vt straight-line trajectories)
and the acceleration induced by the Lorentz force on the charged particle. The
Lorentz force in turn depends on the electromagnetic fields in the plasma. The
Lorentz force due to an electric field accelerates positively charged particles in
the electric field direction, and can trap charged particles in an electric fields
potential well. A quasi-stationary magnetic field causes a charged particle to
execute a cyclotron or Larmor orbit about a magnetic field line. If the magnetic field is inhomogeneous or an electric field is present, there are, in addition,
charged particle drifts in directions perpendicular to the magnetic field direction. Since the overall behavior of a plasma is governed by the sum of what all
its constituent charged particles are doing, in Chapter 4 we investigate the trajectories of charged particles moving in various types of electromagnetic fields.
Having established in Chapters 14 the fundamental processes in plasmas
(collective phenomena, Coulomb collisions, magnetic structure, charged particle
motion), in Chapters 5 and 6 we present the most commonly used descriptions
of plasmas kinetic, two-fluid, and (one-fluid) magnetohydrodynamics (MHD)
and use them to discuss the most fundamental plasma responses to perturbations. Chapter 5 discusses how to obtain the plasma kinetic equation starting
from a microscopic description. Then, various levels of simplified fluid moment
descriptions and approximate plasma responses are deduced e.g., inertial
(fluidlike) for rapid processes and adiabatic for slow processes. Also, general
conditions for stability against growing collective perturbations of the plasma
are noted there. Chapter 6 discusses the main properties of the MHD model
of plasmas equations, equilibrium, Alfven waves and magnetic reconnection
via the small electrical resistivity in a plasma. This chapter also introduces the
important magnetized plasma parameter , which is the ratio of plasma pressure to magnetic energy density, and discusses some of its effects. Finally, both
these chapters conclude with discussions of the types of plasma models that are
used to describe the behavior of both stable and unstable plasmas on various
time and length scales.
REFERENCES AND SUGGESTED READING
The standard introductory level textbook for plasma physics is
Chen, Introduction to Plasma Physics and Controlled Fusion (1974, 84) [?].

Some recently published plasma physics textbooks that are useful complements
or supplements to this standard introductory textbook and this book are

4
Bittencourt, Fundamentals of Plasma Physics (1986) [?].
Chakraboty, Principles of Plasma Mechanics (1978, 90) [?].
Dendy, Plasma Dynamics (1990) [?].
Golant, Zhilinsky and Sakharov, Fundamentals of Plasma Physics (1980) [?].
Goldston and Rutherford, Introduction to Plasma Physics (1995) [?].
Hazeltine and Waelbroeck, The Framework of Plasma Physics (1998) [?].
Nicholson, Introduction to Plasma Theory (1983) [?].
Nishikawa and Wakatani, Plasma Physics, Basic Theory with Fusion Applications (1990) [?].
Schmidt, Physics of High Temperature Plasmas (1966, 79) [?].

Some of the early, influentual textbooks on plasma physics were


Spitzer, Physics of Fully Ionized Gases (1956, 62) [?].
Chandrasekhar, Plasma Physics (1960) [?].
Thompson, An Introduction to Plasma Physics (1962) [?].
Krall and Trivelpiece, Principles Of Plasma Physics (1973) [?].
Longmire, Elementary Plasma Physics (1963) [?].
Arzimovich, Elementary Plasma Physics (1965) [?].

Other textbooks that contain introductory-level discussions of plasma physics


include
Boyd and Sanderson, Plasma Dynamics (1969) [?].
Clemmow and Dougherty, Electrodynamics of Particles and Plasmas (1969) [?].
Hellund, The Plasma State (1961) [?].
Holt and Haskell, Plasma Dynamics (1965) [?].
Ichimaru, Basic Principles of Plasma Physics, A Statistical Approach (1973) [?].
Rosenbluth and Sagdeev, eds., Handbook of Plasma Physics (1983) [?].
Shohet, The Plasma State (1971) [?].
Seshadri, Fundamentals of Plasma Physics (1973) [?].
Tannenbaum, Plasma Physics (1967) [?].

. Useful compendia of plasma physics formulas include:


Book, NRL Plasma Formulary (1977, 1990) [?].
Anders, A Formulary for Plasma Physics (1990) [?].

DRAFT 8:36
August 11, 2003

c
"J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

Chapter 1

Collective Plasma
Phenomena
The properties of a medium are determined by the microscopic processes in it.
In a plasma the microscopic processes are dominated by collective, rather than
binary, charged particle interactions at least for sufficiently long length and
time scales.
When two charged particles are very close together they interact through
their Coulomb electric fields as isolated, individual particles. However, as the
distance between the two particles increases beyond the mean particle separation distance (n1/3 , in which n is the charged particle density), they interact
simultaneously with many nearby charged particles. This produces a collective
interaction. In this regime the Coulomb force from any given charged particle causes all the nearby charges to move, thereby electrically polarizing the
medium. In turn, the nearby charges move collectively to reduce or shield
out the electric field due to any one charged particle, which in the absence of
the shielding decreases as the inverse square of the distance from the particle.
In equilibrium the resultant cloud of polarization charge density around a
charged particle has a collectively determined scale length the Debye shielding length beyond which the electric field due to any given charged particle is
collectively shielded out. That is, the long range force of the Coulomb electric
field is actually limited to a distance of order the Debye length in a plasma.
On length scales longer than the Debye length a plasma responds collectively
to a given charge, charge perturbation, or imposed electric field. The Debye
shielding distance is the maximum scale length over which a plasma can depart
significantly from charge neutrality. Thus, plasmas, which must be larger than a
Debye length in size, are often said to be quasineutral on average electrically
neutral for scale lengths longer than a Debye length, but dominated by the
charge distribution of the discrete charged particles within a Debye length.
Most plasmas are larger than the Debye shielding distance and hence are not
dominated by boundary effects. However, boundary effects become important

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

within a few Debye lengths of a material limiter or wall. This boundary region,
which is called the plasma sheath region, is not quasineutral. Material probes
inserted into plasmas, which are called Langmuir probes after their developer
(in the 1920s) Irving Langmuir, can be biased (relative to the plasma) and
draw currents through their surrounding plasma sheath region. Analysis of
the current-voltage characteristics of such probes can be used to determine the
plasma density and electron temperature.
If the charge density in a quasineutral plasma is perturbed, this induces a
change in the electric field and in the polarization of the plasma. The small
but finite inertia of the charged particles in the plasma cause it to respond
collectively with Debye shielding, and oscillations or waves. When the characteristic frequency of the perturbation is low enough, both the electrons and
the ions can move rapidly compared to the perturbation and their responses are
adiabatic. Then, we obtain the Debye shielding effect discussed in the preceding
paragraphs.
As the characteristic frequency of the perturbations increases, the inertia
of the charged particles becomes important. When the perturbation frequency
exceeds the relevant inertial frequency, we obtain an inertial rather than adiabatic response. Because the ions are much more massive than electrons (the
proton mass is 1836 times that of an electron see Section A.8 in Appendix
A), the characteristic inertial frequency is usually much lower for ions than for
electrons in a plasma. For intermediate frequencies between the characteristic electron and ion inertial frequencies electrons respond adiabatically but
ions have an inertial response, and the overall plasma responds to perturbations
via ion acoustic waves that are analogous to sound waves in a neutral fluid.
For high frequencies above the electron and ion inertial frequencies both
electrons and ions exhibit inertial responses. Then, the plasma responds by
oscillating at a collectively determined frequency called the plasma frequency.
Such space charge oscillations are sometimes called Langmuir oscillations after
Irving Langmuir who first investigated them in the 1920s.
In this chapter we derive the fundamental collective processes in a plasma:
Debye shielding, plasma sheath, plasma oscillations, and ion acoustic waves.
For simplicity, in this chapter we consider only unmagnetized plasmas ones
in which there is no equilibrium magnetic field permeating the plasma. At the
end of the chapter the length and time scales associated with these fundamental
collective processes are used to precisely define the conditions required for being
in the plasma state. Discussions of applications of these fundamental concepts
to various basic plasma phenomena are interspersed throughout the chapter and
in the problems at the end of the chapter.

1.1

Adiabatic Response; Debye Shielding

To derive the Debye shielding length and illustrate its physical significance, we
consider the electrostatic potential around a single, test charged particle in
a plasma. The charged particles in the plasma will be considered to be free
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

charges in a vacuum. Thus, the electrostatic potential in the plasma can be


determined from
E = 2 = q /#0 ,

Poisson equation,

(1.1)

which results from writing the electric field E in terms of the electrostatic potential, E = , in Gausss law see (??) and (??) Section A.2. The charge
density q is composed of two parts: that due to the test charge being considered and that due to the polarization of the plasma caused by the effect of the
test particle on the other charged particles in the plasma. Considering the test
particle of charge qt to be a point charge located at the spatial position xt and
hence representable1 by (x xt ), the charge density can thus be written as
q (x) = qt (x xt ) + pol (x)

(1.2)

in which pol is the polarization charge density.


The polarization charge density results from the responses of the other
charged particles in the plasma to the Coulomb electric field of the test charge.
For slow processes (compared to the inertial time scales to be defined more precisely in Section 1.4 below), the responses are adiabatic. Then, the density of
charged particles (electrons or ions) with charge2 q and temperature T in the
presence of an electrostatic potential (x) is given by [see (??) in Section A.3]
n(x) = n0 eq(x)/T ,

Boltzmann relation (adiabatic response),

(1.3)

where n0 is the average or equilibrium density of these charged particles in the


absence of the potential. The potential energy q of our test particle will be
small compared to its thermal energy, except perhaps quite close to the test
particle. Thus, we expand (1.3) assuming q/T << 1:
n $ n0 (1

q 1 q 2 2
+
),
T
2 T2

perturbed adiabatic response.

(1.4)

The validity of this expansion will be checked a posteriori at the end of this
section. To obtain the desired polarization charge density pol caused by the
effect of the potential on all the charged particles in the plasma, we multiply
(1.4) by the charge q for each species s (s = e, i for electrons, ions) of charged
particles and sum over the species to obtain
#
$%
!
! n0s q 2 "
qs
s
ns q s =
1+O
(1.5)
pol
Ts
Ts
s
s

in which the big oh O indicates the order of the next term in the series
expansion. In obtaining this result we have used the fact that on average a
1 See

Section B.2 in Appendix B for a discussion of the Dirac delta function (x).
this book q will represent the signed charge of a given plasma particle and
e ! 1.602 1019 coulomb will represent the magnitude of the elementary charge. Thus, for
electrons we have qe = e, while for ions of charge Zi we have qi = Zi e.
2 Throughout

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA


plasma is electrically quasineutral:
!

n0s qs = 0,

quasineutrality condition.

(1.6)

Retaining only the lowest order, linear polarization charge density response
in (1.5), substituting it into (1.2), and using the resultant total charge density
in the Poisson equation (1.1), we obtain
$
#
1
qt
(x xt )
(1.7)
2 + 2 =
D
#0
in which the 1/2D term is caused by the polarizability of the plasma. Here, D
is the Debye shielding length:
! 1
! n0s q 2
1
1
n0e e2
n0i Zi2 e2
1
s

= 2 + 2 =
+
,
2
2
D
Ds
#0 Ts
De
Di
#0 Te
#0 Ti
s
s

plasma Debye length.

(1.8)

In the last expressions we have assumed a plasma composed of electrons with


density n0e and only one species of ions with charge Zi e and density n0i . Note
that for comparable electron and ion temperatures the electron and ion Debye
lengths are comparable. The overall plasma Debye length is obtained from the
sum of the inverse squares of the Debye lengths of the various species of charged
particles in the plasma. For a plasma composed of electrons and protons, which
we will call an electron-proton plasma, the lower temperature component will
give the dominant contribution to the overall plasma Debye length. Numerically,
the electron Debye length is given in SI (mks) units by
De

&

#0 Te
$ 7434
ne e2

'

Te (eV)
m,
ne (m3 )

electron Debye length.

(1.9)

The general solution of (1.7) in an infinite, homogeneous three-dimensional


plasma geometry3 is4
t (x) =

qt e|xxt |/D
qt er/D
=
,
{4#0 } |x xt |
{4#0 } r

potential around a test particle.


(1.10)

Here, the subscript t indicates this is the particular solution for the potential
around a test charge qt in a plasma. That this is the solution can be verified by
substituting it into (1.7), noting that (2 (+ 1/2D )t = 0 everywhere
except
((
where r |x xt | 0 and there limr0 d3x 2 = limr0 ! dS =
3 For

one- and two-dimensional geometries see Problems 1.4 and 1.5.


and throughout this book we write the mks factor {4#0 } in braces; eliminating this
factor yields the corresponding cgs (Gaussian) forms for electrostatic response formulas.
4 Here

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

Coul
t
Debye
!!!!!!shielding
bmin n-1/3

Figure 1.1: Potential t around a test particle of charge qt in a plasma and


Coulomb potential Coul , both as a function of radial distance from the test
particle. The shaded region represents the Debye shielding effect. The charac1/3
teristic distances are: D , Debye shielding distance; ne
, mean electron sepcl
2
aration distance; bmin = q /({4#0 }T ), classical distance of closest approach
where the e/T << 1 approximation breaks down.

qt /#0 . The solution given in (1.10) is also the Green function for the equation
(2 + 1/2D ) = free /#0 see Problem 1.6.
The potential around a test charge in a plasma, (1.10), is graphed in Fig. 1.1.
Close to the test particle (i.e., for r |x xt | << D ), the potential is simply the bare Coulomb potential Coul = qt / ({4#0 } |x xt |) around the test
charge qt . For separation distances of order the Debye length D , the exponential factor in (1.10) becomes significant. For separations large compared to
the Debye length the potential t becomes exponentially small and hence is
shielded out by the polarization
cloud surrounding the test charge. Overall,
(
there is no net charge Q V d3x q from the combination of the test charge
and its polarization cloud see Problem 1.7. The difference between t and
the Coulomb potential is due to the collective Debye shielding effect.
We now use the result obtained in (1.10) to check that the expansion (1.4)
was valid. Considering for simplicity a plasma with Ti >> Te [so the electron
Debye length dominates in (1.8)], the ratio of the potential around an electron
test charge to the electron temperature at the mean electron separation distance
1/3
can be written as
of |x xt | = ne
*
+
,1/3 )
3
exp
1/
n

e De
1
et ))
=
$
.
(1.11)
)
3
2/3
1/3
Te |xxt | = ne
4 (ne De )
4 (ne 3De )2/3

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA


For this to be small and validate our expansion in (1.4), we must require
ne 3D >> 1,

necessary condition for the plasma state.

(1.12)

That is, we must have many charged particles (electrons) within a Debye cube
a cube each side of which is the Debye shielding distance in length.5 Physically, (1.12) is a necessary condition for the plasma state because it represents
the requirement that, at the mean interparticle separation distance, collective
interactions of charged particles dominate over binary interactions. The number
of charged particles within a Debye cube (or more often its reciprocal 1/ne 3D )
is called the plasma parameter since it must be large for the medium to be in
the plasma state.
As another check on the validity of the preceding expansion approach, we
next confirm that the electric field energy in the polarization cloud is small
compared to a typical thermal energy for the test particle the temperature
of that species of particles. The polarization electric field is determined by the
difference between the potential t around a test charge in the plasma and the
test charges Coulomb potential Coul :
. +
,/
d q er/D 1
(1.13)
er
Epol = (t Coul ) =
dr
{4#0 }r
er r = (x xt )/|x xt | is a unit vector in the
in which r |x xt | and
r xxt direction. The variation of the polarization electric field as a function
of the distance r away from the test charge is shown in Fig. 1.2.
The energy density associated with this electric field is #0 |Epol |2 /2. Using a
spherical coordinate system whose origin is at the position of the test charge,
we find that the total electric field energy obtained by integrating the energy
density, normalized to the electron temperature (again assuming Ti >> Te for
simplicity) can be written as
1
Te

d3x

#0
|Epol |2
2

$2 0
" # r/D
$%2
#
4#0
d e
1
q
r2 dr
2Te {4#0 }
dr
r
0
I
.
(1.14)
8ne 3D

Here, the dimensionless integral I is simplified using x r/D and is given by


I

"
#
$%2 0 "
%2
d ex 1
1 ex
dx x
=
dx ex
dx
x
x
0
0
/ 0
0 .
2

x
, (1 e )
2 + x
1
e e2x +
=
dx e2x
dx e2x =
=
2
x
x
2
0
0

5 Since the intrinsic geometry of the polarization cloud around a test charge is spherical,
plasma physicists often use as the appropriate measure the number of charged particles within
a Debye sphere, (4/3)ne 3D , which by (1.12) must also be large compared to unity.

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

Er

Epol

Figure 1.2: Coulomb and polarization radial electric fields around a test particle of charge q in a plasma. Because the plasma polarization acts to shield
out the positive Coulomb electric field, the polarization electric field is negative.
The polarization electric field is finite at the origin, decays smoothly with distance away from the test charge, and shields out the Coulomb electric field for
separations larger than the Debye length D .

in which in the first integral form on the second line we have integrated the last
term by parts and cancelled it with the second term, and the final integral is
evaluated using (??) in Appendix C. From (1.14) we again see that our expansion
approach is valid as long as there are many electrons within a Debye cube (or
sphere), since then the electric field energy in the polarization cloud around a
test charge is small compared to the typical, thermal energy of a charged particle
in a plasma.
We can also use the concepts developed in the preceding discussion to estimate the level of thermal fluctuations or noise in a plasma. The thermal
fluctuations are caused by the interactions between charged particles through
the electric field around one particle influencing the positions of other particles within approximately a Debye sphere around the original charged particle.
That is, they are caused by correlations between particles, or by electric field
correlations within the plasma. To calculate these properly requires a plasma
kinetic theory (see Chapter 13). However, the fluctuation level can be estimated
as follows.
A relevant measure of the magnitude of the thermal noise in a plasma is the
2 /2 to the therratio of the electric field energy density in the fluctuations #0 |E|
mal energy density nT . The polarization electric field given by (1.13) represents
the correlation electric field between a test particle at xt and an observation
point x. From Fig. 1.2 we see that the polarization electric field is localized to
within about a Debye length of any given charge, and its magnitude there is
not too different from its value at r |x xt | = 0: Epol (0) = q/(2{4#0 }2D ).
Also, we note that all charged particles within about a Debye sphere [namely

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

(4/3)ne 3D particles] will contribute to the electric field fluctuations at any


given point. Hence, omitting numerical factors, we deduce that the scaling of
the relative electric field fluctuation energy from two-particle correlations in
a plasma is given by
#
$*
4
#0
2
3
#0 2
n

|E
(0)|
e
pol
D
|E|
1
3
2
2

<< 1,
ne T e
ne Te
ne 3D
thermal fluctuation level.
(1.15)
We thus see that, as long as (1.12) is satisfied, the thermal fluctuation level is
small compared to the thermal energy density in the plasma and again our basic
expansion approach is valid. The thermal fluctuations occur predominantly at
scale-lengths of order the Debye length or smaller. The appropriate numerical
factor to be used in this formula, and the frequency and wave-number dependence of the thermal fluctuations in a plasma, can be obtained from plasma
kinetic theory. They will be discussed and determined in Chapter 13.

1.2

Boundary Conditions; Plasma Sheath

A plasma should be larger than the Debye shielding distance in order not to
be dominated by boundary effects. However, at the edge of a plasma where
it comes into contact with a solid material (e.g., a wall, the earth), boundary
effects become important. The region where the transformation from the plasma
state to the solid state takes place is called the plasma sheath.
The role of a plasma sheath can be understood as follows. First, note that
for comparable electron and ion temperatures the typical1electron speed, which
will be taken to be the electron thermal speed vT e 2Te /me [see (??) in
Section
A.3], is much larger than the typical (thermal) ion speed (vT e /vT i
1
> 43 >> 1). Since the electrons typically move much faster than
mi /me
6
the ions, electrons tend to leave a plasma much more rapidly than ions. This
causes the plasma to become positively charged and build up an equilibrium
electrostatic potential that is large enough [ a few Te /e, see (1.23) below]
to reduce the electron loss rate to the ion loss rate so the plasma can be
quasineutral in steady state. The potential variation is mostly localized to the
plasma sheath region, which is of order a few Debye lengths in width because
that is the scale length on which significant departures from charge neutrality
are allowed in a plasma. Thus, a plasma in contact with a grounded wall will:
charge up positively, and be quasineutral throughout most of the plasma, but
have a positively charged plasma sheath region near the wall.
We now make these concepts more concrete and quantitative by estimating
the properties of a one-dimensional sheath next to a grounded wall using a simple
plasma model. Figure 1.3 shows the specific geometry to be considered along
6 Many people use an analogy to remember that electrons have much larger thermal velocities than ions: electrons are like fast moving, lightweight ping pong balls while ions are like
slow-moving, more massive billiard balls for equal excitation or thermal energies.

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

plasma
sheath

presheath

bulk!plsama

n
n

xs

ion!rich
region

transition

quasi-neutral!plasma

ni
ne

xs

Figure 1.3: Behavior of the electrostatic potential and electron and ion densities
in the sheath, presheath (or transition) and bulk plasma regions of a plasma
2
= 0.9. The
in contact with a grounded wall. For the case shown Te /mi V
sheath parameters determined in the text are $ 3 Te /e and xS $ 2 De .
The long-dash line in the top figure indicates the approximation given in (1.26).

with the behavior of the potential, and electron and ion densities in the plasma
sheath and bulk plasma regions, as well as in the presheath (or transition) region
between them.
The electron density is determined from the Boltzmann relation (1.3):
2
3
e [(x) ]
(1.16)
ne (x) = n exp
Te
in which (x) indicates the equilibrium potential profile in the plasma, and the
subscript indicates evaluation of the quantities in the bulk plasma region
far from the wall (i.e., beyond the plasma sheath and presheath regions whose
properties we will determine). (In using this equation it is implicitly assumed
that the background electron velocity distribution is Maxwellian.)
For simplicity, we consider an electron-proton plasma with negligible ion
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

10

thermal motion effects (e >> Ti ). The potential variation in and near the
sheath produces an electric field that increases the flow of ions toward the wall,
which will be assumed to be grounded. The ion flow speed Vi in the x direction
is governed by conservation of energy for the cold (e >> Ti ) ions:
1
1
2
mi Vi2 (x) + e(x) = constant = mi V
+ e
(1.17)
2
2
in which we have allowed for a flow of ions from the bulk of the plasma into the
presheath region so as to ultimately balance the electron flow to the wall. The
ion flow at any given point is given by
'
&
%
"
2
2e
2e
mi V
2
(x) .
[ (x)] =
+
Vi (x) = V +
mi
mi
2e
The spatial change in the ion flow speed causes the ion density to change as
well a high flow speed produces a low ion density. The ion density variation
is governed, in a steady equilibrium, by the continuity or density conservation
equation [see (??) in Appendix A] for the ion density: d(ni Vi )/dx = 0, or
ni (x)Vi (x) = constant. Thus, referencing the ion density to its value n in the
bulk plasma (x ), it can be written as
2
31/2
2e [ (x)]
.
(1.18)
ni (x) = n 1 +
2
mi V

Substituting the electron and ion densities into Poissons equation (1.1), we
obtain the equation that governs the spatial variation of the potential in the
sheath, presheath and plasma regions:
e
d2
= (ni ne )
dx2
#0
.2
2
31/2
3/
e [ (x)]
2e [ (x)]
n e
exp
1+
.(1.19)
=
2
#0
mi V
Te

While numerical solutions of this equation can be obtained, no analytic solution


is available. However, limiting forms of the solution can be obtained near the
wall (x << xS ) and in the bulk plasma (x >> xS ). Even though a simple
solution is not available in the transition region, solutions outside this region
can be used to define the sheath position xS and the conditions needed for
proper sheath formation.
In the quasineutral plasma far from the plasma sheath region (x >> xS ) the
potential (x) is very close to its asymptotic value . In this region we approx2
<<1 and
imate the electron and ion densities in the limits 2e [ (x)] /mi V
e [ (x)] /Te <<1, respectively:
3
2
e [ (x)]
+ ,
ne (x) $ n 1
Te
2
3
e [ (x)]
ni (x) $ n 1
+

.
2
mi V
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

11

Keeping only linear terms in (x), (1.19) can thus be simplified to


#
$
1
d2 [ (x)]
Te
$
(1.20)
1

[ (x)] .
2
dx2
2De
mi V
Here, De is the electron Debye length evaluated at the bulk plasma density n .
2
For mi V
< Te the coefficient of (x) on the right would be negative;
this would imply a spatially oscillatory solution that is not physically realistic for the present plasma model, which implicitly assumes that the potential
is a monotonic function of x. Thus, a necessary condition for proper sheath
formation in this model is
|V |

1
Te /mi ,

Bohm sheath criterion.

(1.21)

Since this condition need only be satisfied marginally and the ion flow into
the sheath region typically assumes its minimum value, it is usually sufficient
to make this criterion an equality. The Bohm sheath criterion implies that ions
must enter the sheath region sufficiently rapidly to compensate for the electron
charge leakage through the sheath to the wall. In general, what is required for
proper sheath formation is that, as we move toward the wall, the local charge
density increases as the potential decreases:
1 q / < 0 for all x. Also, since
we will later find (see Section 1.4) that Te /mi is the speed of ion acoustic
waves in a plasma (for the plasma model being considered), the Bohm sheath
criterion implies that the ions must enter the presheath region at a supersonic
speed relative to the ion acoustic speed.
As long as the Bohm sheath criterion is satisfied, solutions of (1.19) will be
well-behaved, and exponentially damped in the presheath region: for x >> xS
2 1/2
we have (x) $ C exp(x/h) where h = De (1 Te /mi V
)
and C is
a constant of order S . Thus, for1this plasma model, in the typical case
where V is equal to or slightly exceeds Te /mi , the presheath region where the
potential deviates from extends only a few Debye lengths into the plasma. In
more comprehensive models for the plasma, and in particular when ion thermal
effects are included, it is found that the presheath region can be larger and the
potential variation in this region is influenced by the effects of sheath geometry,
local plasma sources, collisions and a magnetic field (if present). However, the
Bohm sheath criterion given by (1.21) remains unchanged for most physically
relevant situations, as long as the quantity on the right side is interpreted to be
the ion acoustic speed in the plasma model being used.
We next calculate the plasma potential that the plasma will rise to in
order to hold back the electrons so that their loss rate will be equal to the ion loss
rate from the plasma. The flux of ions to the wall is given by ni V , which
when1evaluated at the Bohm sheath criterion value given in (1.21) becomes
n Te /mi . (The flux is negative because it is in the negative x direction.)
A Maxwellian distribution of electrons will produce (see Section A.3) a random
flux of electrons to the 1
wall on the left side of the plasma of (1/4)ne ve =
(n /4) exp(e /Te ) 8Te /me . Thus, the net electrical current density to
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

12

the wall will be given by


J = Ji Je = e(ni V ne ve )
*1
1
= en
Te /mi (1/4) 8Te /me exp (e /Te ) .

(1.22)

Since in a quasineutral plasma equilibrium we must have J = 0, the plasma


potential is given in this plasma model by

Te
=
ln
e

&

Te
Te
mi
2.84
$ 3 ,
2me
e
e

plasma potential,

(1.23)

where after the inequality we have used the proton to electron mass ratio mi /me
= 1836. In the original work in this area in 1949, Bohm argued that a potential
drop of Te /2e extending over a long distance into the plasma (much further
than where
we are calculating) is required to produce the incoming ion speed
1
V Te /mi at the sheath edge. In Bohms model the density n is e1/2 =
0.61 times smaller than the bulk plasma ion density and thus the ion current
Ji is smaller by this same factor. For this model, the potential in (1.23)
increases by 0.5 Te /e to 3.34 Te /e. Since lots additional physics (see end of
preceding paragraph) needs to be included to precisely determine the plasma
potential for a particular situation, and the plasma potential does not change
too much with these effects, for simplicity we will take the plasma potential
to be approximately 3 Te /e.
Finally, we investigate the form of (x) in the sheath region near the wall
(x << xS ). In this region the potential is much less than and the electron
density becomes so small relative to the ion density that it can be neglected. The
equation governing the potential in this ion-rich region can thus be simplified
from (1.19) to
%1/2
"
2
en
/2e
mi V
d2 (x)
$
.
2 /2e (x)
dx2
#0
+ mi V

(1.24)

This equation can be de-dimensionalized by multiplying through by e/Te . Thus,


defining a dimensionless potential variable by
5
4
2
/2e (x)
e + mi V
,
(1.25)
(x)
Te
the equation can be written as
1
d2
$ 2 .
dx2

2
/2Te )1/4 .
in which De /(mi V
To integrate this equation we multiply by d/dx and integrate over x us

ing (d/dx)(d2 /dx2 ) = (1/2)(d/dx)(d/dx)2 and dx(d/dx)/ = d/ =

2d to obtain

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA


1
2

d
dx

$2

13

2
+ constant.
2

Since both and d/dx are small near xS , this equation is approximately valid
for the x < xS region if we take the constant in it to be zero. Solving the
resultant equation for d/dx, we obtain
21/4
d
$
dx

4
2dx
d(3/4 ) $
.
3

2
Integrating this equation from x = 0 where = 0 (e + mi V
/2)/Te to
x where = (x), we obtain
3/4

3/4 (x) 0

3x
,
2

or
2
(x) $ ( + mi V
/2e)[1 (1 x/xS )4/3 ].

(1.26)

Here, we have defined


3/4

20
xS
3

25/4
=
3

Te
2
mi V

$1/4 #

$3/4
2
/2
e + mi V
De ,
Te
sheath thickness.

(1.27)

Equation (1.26) is valid in the sheath region near the wall (0 < x << xS ). We
have identified the scale length in (1.27) with the sheath width xS because this
is the distance from the wall at which the potential (x) extrapolates to the
2
/2e.
effective plasma potential in the bulk plasma, + mi V
1
Te /mi , the sheath
Using the value for given in (1.23) and V $
thickness becomes xS $ 2 De . Thus, as shown in Fig. 1.3, for this model the
plasma charges to a positive potential of a few Te /e and is quasineutral up
to the non-neutral plasma sheath region, which extends a few Debye lengths
( 2 xS 4 De in Fig. 1.3) from the grounded wall into the plasma region.

1.3

Langmuir Probe Characteristics*

To further illuminate the electrical properties of a static or equilibrium plasma,


we next determine the current that will be drawn out of a probe inserted into an
infinite plasma and biased to a voltage or potential B . Such probes provided
some of the earliest means of diagnosing plasmas and are called Langmuir probes,
after Irving Langmuir who developed much of the original understanding of
their operation. The specific situation to be considered is sketched in Fig. 1.4.
For simplicity we assume that the probe is small compared to the size of the
plasma and does not significantly disturb it. The probe will be assumed to
have a metallic (e.g., molybdenum) tip and be electrically connected to the
outside world via an insulated tube through the plasma. Probes of this type are
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

14

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

ISe

plasma

ISi

Figure 1.4: Schematic of Langmuir probe inserted into a plasma and its idealized
current-voltage characteristics: current I drawn out of the probe as a function
of the bias voltage or potential B . The labeled potentials and currents are:
f , floating potential; p , plasma potential; ISi , ion saturation current; ISe ,
electron saturation current.

< 10 eV,
often used in laboratory plasmas that have modest parameters (Te
19
3
<
ne 10 m probes tend to get burned up at higher plasma parameters).
Since the bias potential B on the probe will not affect the incoming ion
flow speed V (for B < ), following the discussion leading to (1.22) the ion
current out of the probe will be given by
1
Ii = AS Ji = n e Te /mi AS ISi , ion saturation current (ISi ), (1.28)
where AS is the area of the probe plus sheath over which the ions are collected
by the probe. For the electrons we must take account of the bias potential B
on the probe. The electron current into the probe is given by

1
n e Te /2me Ap ISe ,
B p ,
Ie = Ap Je =
n e1T /2m A exp [ e( )/T ] , < ,

e
e p
p
B
e
B
p
(1.29)

in which Ap is the area of the probe and p is the plasma potential the voltage
at which all electrons heading toward the probe are collected by it. (Whereas
the effective area for ions to be collected by the probe encompasses both the
probe and the sheath, for B < p the relevant area Ap for electrons is just the
probe area since only those electrons surmounting the sheath potential make
it to the probe see Fig. 1.3. However, when B > p the relevant area,
and consequent electron current, grows slightly and roughly linearly with bias
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

15

voltage, which is then attracting electrons and modifying their trajectories in


the vicinity of the probe. In the idealized current-voltage curve in Fig. 1.4 we
have neglected this latter effect.)
The total current I = Ii + Ie drawn to the probe is shown in Fig. 1.4
as a function of the bias voltage or potential B . For a large negative bias
the electron current becomes negligible and the current is totally given by the
ion current ISi , which is called the ion saturation current. The potential f
at which the current from the probe vanishes is called the floating potential ,
which is zero for our simple model. However, it is often slightly negative in real
plasmas, unless there is secondary electron emission from the probe, in which
case it can become positive. For potentials larger than the plasma potential
p all electrons on trajectories that intercept the probe are collected and the
current is given by the electron saturation current ISe . Except for differences
in the charged particle collection geometry (typically cylindrical or spherical
probes versus a planar wall), in the sheath thickness (relative to probe size)
effects and perhaps in secondary electron emission, the difference between the
plasma and floating potentials is just the naturally positive plasma potential
that we derived in (1.23). That is, p f $ 3 Te /e.
In a real plasma the idealized current-voltage characteristic that is indicated
in Fig. 1.4 gets rounded off and distorted somewhat due to effects such as charged
particle orbit effects in the sheath, probe geometry, secondary electron emission
from the probe and other effects. Indeed, because of the practical importance of
Langmuir probes in measuring plasma parameters in many laboratory plasmas,
as we will discuss in the next paragraph, there is a large literature on the
current-voltage characteristics of various types of probes in real plasmas (see
references and suggested reading at the end of the chapter). Nonetheless, the
basic characteristics are as indicated in Fig. 1.4.
For bias potentials that lie between the floating and plasma potentials, the
current from the probe increases exponentially with bias potential B . Thus,
the electron temperature can be deduced from the rate of exponential growth in
the current as the bias potential is increased: Te /e $(I ISi )/(dI/dB ). Alternatively, one can use a double probe to determine the electron temperature
see Problem 1.11. If the electron temperature is known, the plasma ion
1 density
can be estimated from the ion saturation current: n $ ISi /(eAS Te /Mi ).
Langmuir probes are thus important diagnostics for measuring the plasma density and electron temperature in laboratory plasmas with modest parameters.
The thickness of the plasma sheath changes as the bias potential B is
varied. The derivation of the sheath thickness xS given in (1.24) to (1.27) can
be modified to account for the present biased probe situation by replacing the
2
potential with p B . Thus, setting mi V
/Te to unity to satisfy the
Bohm sheath criterion (1.21), the sheath thickness around a biased probe in a
plasma is given approximately by
xS $

25/4
3

DRAFT 10:26
August 12, 2003

p + 0.5 B
Te /e

$3/4

De ,

c
!J.D
Callen,

sheath thickness.

(1.30)

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

16

This formula is valid for e(p +0.5B ) >> Te small or negative bias voltages
B p . As the bias potential B increases toward the plasma potential p ,
< 0.5 Te .
the plasma sheath becomes thinner; it disappears for e(p B )
For large negative bias potentials (|B | >> Te /e), the electrical current
density flowing through the ion-rich sheath region is limited by space charge
effects and given by the Child-Langmuir law see Problem 1.13. However, transiently the current density can be larger that indicated by the Child-Langmuir
law see Problem 1.14.

1.4

Inertial Response; Plasma Oscillations

In the preceding sections on Debye shielding and its effects we considered the
adiabatic or static response of charged particles and a plasma to the Coulomb
electric field around a charged particle in the plasma. Next, we discuss the
inertial (or dynamic) response of a plasma. To do this we consider the electric
polarization response of charged particles and a plasma to a small electric field
perturbation, which may be externally imposed or be collectively generated
within the plasma.
First, we calculate the motion of a charged particle in response to an electric
field. The velocity v of a charged particle of mass m and charge q subjected

t) is governed by Newtons second law


to an electric field perturbation7 E(x,

(F = ma) with force q E:


m

dv

= q E(x,
t).
dt

(1.31)

The electric field perturbation will be assumed to be small enough and sufficiently slowly varying in space so that nonlinear and translational motion effects
are negligible. Thus, it will be sufficient to evaluate the electric field at the initial position x0 and neglect the small variation in the electric field induced by
the motion x(t) of the charged particle. This approximation will be discussed
further after the next paragraph.
(t) induced by the
Integrating (1.31) over time, the velocity perturbation v
electric field perturbation for a particle with initial velocity v0 is given by
0 t
0 t
q
$ (t$ ), t$ ] $ q
0 , t$ ).
(t) v(t) v0 =
dt$ E[x
dt$ E(x
(1.32)
v
m 0
m 0

Integrating once more over time, we find that the motion induced by the electric
field perturbation becomes
0 t
0 t#
q
0 , t$$ ), inertial response.
(t) x(t) (x0 + v0 t) $
dt$
dt$$ E(x
x
m 0
0
(1.33)
7 Perturbations to an equilibrium will be indicated throughout the book by a tilde over
the symbol for the perturbed quantity. Equilibrium quantities will be indicated by 0 (zero)
subscripts.

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

17

Because the response of the particle to the electric field force is limited by the
inertial force m a = m dv/dt, this is called an inertial response. Note that this
response is inversely proportional to the mass of the charged particle; thus,
the lighter electrons will give the primary inertial response to an electric field
perturbation in a plasma.
We check our approximation of evaluating the electric field at the initial
position x0 by expanding the electric field in a Taylor series expansion about
the charged particle trajectory given by
|x +
[x(t), t] = E
(x0 , t) + (
x + v0 t) E
E
0

(1.34)

Our approximation is valid as long as the second (and higher order) terms in
this expansion are small compared to the first term:
<< E.

(
x + v0 t) E

(1.35)

Thus, the electric field perturbation must vary sufficiently slowly in space (i.e.,

1 must be long compared to the distance


the gradient scale length |(1/|E|)
E|
is small compared to
E
|
x + v0 t|), be small enough (so the nonlinear term x

E) and the elapsed time must not be too long. These approximations will be
checked a posteriori at the end of this section.
of a charged particle in response to the electric field
The inertial motion x
perturbation creates an electric dipole moment q
x. A uniform density n0 of such
= n0 q
x. Summing
charged particles leads to an electric polarization density P
over the species of charged particles in the plasma, the total plasma polarization
density becomes
=
P

s = #0
n0s qs x

2
ps

dt$

t#

(t$$ )
dt$$ E

(1.36)

in which for each charged species s


2

ps

n0s qs2
,
ms #0

square of species plasma frequency,

(1.37)

is the inertial or plasma frequency for a species s, whose physical significance


will be discussed below.
Because the ions are so much more massive than the electrons (the ratio of
the proton to electron mass is 1836), they have much more inertia. Thus, their
plasma frequency is much
1 smaller than that for the electrons for example,
for protons pi /pe = me /mp $ 1/43 << 1. Since the electrons give the
dominant contribution to the plasma polarization and have the largest plasma
frequency, we have
!
2
2
2
2
ps
= pe
+ pi
$ pe
.
(1.38)
s

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

18

Numerically, the electron plasma frequency is given by


pe
or

'

1
ne e2
$ 56 ne (m3 ) rad/sec, radian plasma frequency,
me #0
1
fpe pe /2 $ 9 ne (m3 ) Hz,

plasma frequency.

(1.39)

(1.40)

The plasma polarization in (1.36) causes [see(??) and (??) Section A.2] a polarization charge density pol given by the negative of the divergence of the

polarization P:
0 t
0 t#
!
2
$

$$ ).
ps
dt
dt$$ E(t
(1.41)
pol (E) = P = #0
s

in a plasma we need to use


Now, to calculate the perturbed electric field E
Gausss law, which is given in (1.1). For the charge density q we imagine that
there are polarization charge densities due to both the electric field perturbation
we have been considering, and an externally imposed electric field Eext which
namely condition (1.35). Thus, the relevant
satisfies the same conditions as E
form of Gausss law becomes
0 t
0 t#
*
!
1
2
$
$$ ) + Eext (t$$ ) .

ps
dt
dt$$ E(t
(1.42)
E = pol =
#0
0
0
s
This differential and integral equation in space and time, respectively, can be
reduced to a simpler, completely differential form by taking its second partial
derivative with respect to time to yield
/
.
9
:
!
2E
2

+
ps E + Eext
= 0.
(1.43)

t2
s

Using the approximation in (1.38), we thus find that taking into account the
inertial effects of charged particles (mostly electrons), nontrivial (i.e., nonvanishing) electric field perturbations satisfying condition (1.35) are governed by
the differential equation

2E
2
2
+ pe
Eext .
E = pe
2
t

(1.44)

This is a linear, inhomogeneous differential equation of the harmonic oscilla induced by the
tor type with frequency pe for the perturbed electric field E
externally applied electric field Eext .
The complementary (in the current langauge of mathematics) solutions of
the homogeneous part of this equation are of the form
h = Cc cos pe t + Cs sin pe t,
E
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

(1.45)

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

19

plasma

V(t)

Figure 1.5: Schematic of circuit for imposing an oscillating potential (t) =


0 sin 0 t across a plasma.

where Cc and Cs are arbitrary coefficient vectors to be fixed by the boundary


conditions. These plasma oscillation solutions show that the plasma responds
inertially to electric field perturbations by oscillating at the electron plasma
frequency pe . Externally imposed electric fields will induce perturbations in
the plasma that are combinations of the time dependence of the externally
imposed field and the electron plasma oscillations.
In the present simple model plasma oscillations are undamped. Collisions
(electron-neutral or Coulomb) damp them at rates proportional to the relevant
collision frequency see Problem 1.18. Also, as we will discuss in Chapter 8,
kinetic effects will lead to evanescence of these oscillations due to wave-particle
resonance effects Landau damping.
To illustrate the plasma responses more concretely, we consider the response
of a plasma to an externally imposed sinusoidal electric field. (An alternative
illustration for just plasma oscillations is developed in Problem 1.19 using a
one-dimensional plasma slab model.) As shown in Fig. 1.5, the electric field will
be induced by imposing an oscillating potential (t) = 0 sin 0 t at time t = 0
across plates on opposite sides of a plasma of thickness L (implicitly >> D )
in the x direction. For simplicity the plasma will be assumed to be infinite in
extent (or at least >> L) in the other two directions so that their effects can
be neglected. Thus, the applied electric field will be given for t > 0 by
Eext =

ex sin 0 t E0 sin 0 t.
L

(1.46)

The particular solution of (1.44) in response to this externally applied electric

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

20

field is
p =
E

2
pe
E0 sin 0 t.
2
02 pe

(1.47)

Adding together the homogeneous, particular and externally applied electric


h +E
p + Eext ) of the solution of (1.44), and
+ Eext = E
field components (E = E
subjecting them to the boundary conditions that E(t =4 0) = 0 +and dE/dt
,5|t=0
2
= dEext /dt |t=0 = 0 E0 , we find Cc = 0 and Cs = 0 pe / 02 pe
E0 .
Hence, the total electric field E driven by Eext is given for t > 0 by
E(t)

2
pe
0 pe
E
sin

t
+
E0 sin 0 t + E0 sin 0 t
0
pe
2
2
2
2
0 pe
0 pe

02
0 pe
E
sin

t
+
E0 sin 0 t
0
pe
2
2
02 pe
02 pe
driven sin 0 t.
plasma sin pe t + E
E

(1.48)

driven oscillating at
The frequency dependences of the net driven response E
plasma oscillating at the plasma frequency
frequency 0 and of the response E
pe are shown in Fig. 1.6. For 0 much less than the electron inertial or plasma
driven is of order 2 / 2 compared with the
frequency pe , we find that E
pe
0
externally applied electric field E0 sin 0 t, and hence tends to be small. In
this limit the electrons have little inertia (0 << pe ) and they develop a
strong polarization response that tends to collectively shield out the externally
applied electric field from the bulk of the plasma. In the opposite limit 02 >>
2
, the inertia of the electrons prevents them from responding significantly,
pe
their polarization response is small, and the externally imposed electric field
<< Eext . The singularity
permeates the plasma in this limit E $ Eext since E
at 0 = pe indicates that when the driving frequency 0 coincides with the
natural plasma oscillation frequency pe the linear response is unbounded. In
Chapters 7 and 8 we will see that collisions or kinetic effects bound this response
and lead to weak damping effects for 0 $ pe . Nonlinear effects can also lead
to bounds on this response.
plasma response in (1.48), which oscillates at the plasma frequency, is
The E
caused by the electron inertia effects during the initial turn-on of the external
electric field. Note that it vanishes in both the low and high frequency limits
because for low 0 the excitation is small for the nearly adiabatic (0 << pe )
turn-on phase, while for high 0 the electron inertial response is small during the
very brief (t 1/0 << 1/pe ) turn-on phase. Like the driven response, the
plasma response becomes unbounded in this simple plasma model for 0 pe .
Finally, we go back and determine the conditions under which the approximation (1.35) that we made in calculating the plasma polarization induced by
an electric field is valid. Referring to the physical situation shown in Fig. 1.5, we

1
take the gradient scale length of the electric field perturbation |(1/|E|)
E|
to be of order the spacing L between the plates. We first estimate the condition
imposed by the particle streaming indicated by the term v0 t in (1.35). To make
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

21

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA


collective
shielding

dielectric
medium

E0
0

wpe

Figure 1.6: Frequency dependence of the electric field components oscillat driven , solid lines) and the plasma frequency
ing at the driven frequency 0 (E
plasma , dashed lines) induced in a plasma by Eext = E0 sin 0 t, as indipe (E
cated in Fig. 1.5. The driven frequency response is shielded out for 0 << pe ;
it approaches the imposed electric field for 0 >> pe . The plasma frequency
response is induced by the process of turning on the external electric field; it
becomes small when 0 is very different from pe . The singular behavior for
0 pe results from driving the system at the natural oscillation frequency of
the plasma, the plasma frequency; it is limited in more complete plasma models
by collisional, kinetic or nonlinear effects.

this estimate
we take v0 to be of order the most probable electron thermal speed
1
vT e 2Te /me [see (??) in Section A.3]. Also, we estimate t by 1/. However,
since the most important plasma effectsoccur for pe (see Fig. 1.6), we
scale to pe . Then, since vT e /pe = 2 De , the particle streaming part of
(1.35) leads, neglecting numerical factors, to the condition
L >> De (pe /) .

(1.49)

That is, for pe the plasma must be large compared to the electron Debye
length.
<< E
we consider a situation
E
For validity of the nonlinear condition x

where E = (/L) sin t. Then, again neglecting numerical factors, we find that
to neglect the nonlinearities we must require
# 2$

L2

e
<< 2
.
(1.50)
2
Te
De pe
Since we can anticipate from physical considerations that potential fluctuations
are at most of order some modest factor times the electron temperature in a

plasma, the nonlinear criterion is usually well satisfied as long as the streaming
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

22

criterion in (1.49) is. Hence, our derivation of the plasma polarization is generally valid for pe plasma oscillation phenomena as long as the plasma
under consideration is much larger than the electron Debye length.
We can also use the preceding logic to specify the temporal and spatial
scales on which the inertial response and effects discussed in this section apply
in an infinite, homogeneous plasma versus the conditions where the adiabatic
response in the first section of this chapter apply. (For a general discussion of
inertial and adiabatic responses for a harmonic oscillator see Appendix E.)
For pe , as long as the scale length L x of interest is long compared to the
electron Debye length De , conditions (1.35), (1.49) and (1.50) are all satisfied.
Then, the inertial and electron plasma oscillation effects we have discussed are
relevant since >> vT e /x, which is the inverse of the time required for a
thermal electron to move a distance x. However, for low frequencies << pe
such that x << De (/pe ), or for scale lengths x << De with pe , the
inequality conditions become reversed and the approximations we have used
in this section break down. Then, instead of an inertial response, we obtain
an adiabatic response for << vT e /x and the Debye shielding effects we
discussed in the first section of this chapter. Intermediate situations with x
De (pe /) vT e / must be treated kinetically see Chapter 8.

1.5

Plasma as a Dielectric Medium

is comIn general, any vector field such as the electric field perturbation E
,= 0) and transverse (solenoidal,
posed of both longitudinal (irrotational, E
= 0) parts see Section D.5 of Appendix D. From the form of (1.43) it
E
is clear that we have been discussing the longitudinal component of the electric

= ,
field perturbation. This component is derivable from a potential, E
and represents the electrostatic component of the electric field perturbation.
= 2 ,= 0, we see from Gausss law (1.1) that these elecSince we have E
trostatic perturbation components correspond to charge density perturbations
in the plasma. Thus, the electron plasma oscillations we have been discussing
are electrostatic space charge oscillations in which the longitudinal component
of the electric field and plasma polarization oscillate out of phase with respect
2
2
2

= 2 ( P)/t

= pe
E
.
to each other, i.e., 2 ( E)/t
The polarizability of the plasma by an electric field perturbation can also be
interpreted by considering the plasma to be a dielectric medium. To illustrate
this viewpoint, we note that in a dielectric medium Gausss law becomes [see
(??) in Section A.2]
D = free ,

(1.51)

D #E

(1.52)

where
is the displacement vector, free is the charge density of the free charges (i.e.,
those not contributing to the plasma dielectric), and # is the dielectric constant of
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

23

induces the
the medium (# = #0 for a vacuum). The electric field perturbation E

polarization charge density given in (1.41) and the polarization P. Comparing


(1.51) with (1.41) and (1.42), we deduce that the perturbed displacement vector
is related to the polarization P
through [see (??) and (??)]
D
+P
#0 (1 +
# E,

= #0 E
E ) E
D

(1.53)

#0
E E
P

(1.54)

with

in which
E is the electric susceptibility of the plasma. We have placed hats
over # and E to emphasize that these quantities are only defined with respect
to temporally (and later spatially) varying electric fields; that is, unlike regular dielectric media, their static, homogeneous plasma limits are divergent and
hence do not exist (see below).
=E
sin t that
For the sinusoidal electric field perturbations of the form E
given by (1.36) becomes
we have been discussing, the polarization density P
= #0
P
hence, we have

2
! ps

E () =
and
#I () = #0

2
! ps
s

<

$ #0

#0

E E;
E

2
! ps
s

2
pe
1 2

$
<

2
pe
2

inertial dielectric.

(1.55)

(1.56)

(1.57)

we have performed the integrals in (1.36) as inIn obtaining this form of P


definite integrals in time and hence neglected the initial conditions because
in determining dielectric properties of a medium one considers only the time
asymptotic response and neglects the initial transient effects.
The frequency dependence of the inertial dielectric8 #I () in (1.57), which
represents the inertial response of a plasma, is shown in Fig. 1.7. The fact
driven component in (1.48)
that #I () #0 for >> pe shows why the E
approaches the externally applied electric field in this vacuum limit. Since
#I () is negative for < pe , the externally applied electric field is shielded
8 For media such as water the dielectric response function is nearly constant over most relevant frequency ranges, e.g., for visible light. Hence its properties are characterized by a dielectric constant. However, in plasmas the dielectric response function often varies significantly
with frequency (and wavenumber k). Thus, in plasmas we will usually try to avoid speaking
of a dielectric constant; instead we will just refer to the plasma dielectric. However, when
the dielectric response function is evaluated for a particular frequency (and wavenumber k),
we will often call it the dielectric constant.

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

24

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

"0

wpe

Figure 1.7: Frequency dependence of inertial response plasma dielectric.

out of the plasma or cut off in this frequency range. The vanishing of #I for
= pe indicates that this is a normal mode of oscillation of the plasma, as
is evident from the plasma oscillator equation (1.44) driven electric fields at
frequencies where the dielectric vanishes lead to unbounded resonant responses
in linear theory, as can be inferred from (1.51) and (1.52). Also, since #I is small
plasma is largest
for close to pe , the transient plasma frequency response E
in this frequency range. Finally, we note that #I () is divergent in the 0
or static limit. Thus, the inertial dielectric response of a plasma can only be
defined for temporally varying processes.
Because the polarization of the plasma is 180 out of phase with respect
to the electric field perturbations for all real , the inertial plasma response is
reactive (i.e., not dissipative) for all frequencies . That there is no dissipation
J

can be demonstrated explicitly by calculating the average Joule heating E

v over an oscillation period 2/ and showing that it vanishes.


with J = n0 e
If dissipative effects, such as collisons, are added, they lead to wave damping
through the joule heating they induce in the plasma see Problem 1.18
The energy density of plasma oscillations is composed of two parts: the
2 /2 and the polarization energy denvacuum electric field energy density #0 |E|
1
1
2

E |E| . For an electric field perturbation E


sity wpol = 2 P E = 2 #0
oscillating at frequency the polarization is given in (1.55). Thus, we find
2
2 . Hence, the total energy density [see (??)] in an
wpol = (#0 /2)(pe
/ 2 )|E|
electrostatic plasma oscillation is given by
;
<
2
pe
#0 2
1
#0
2 , wave energy density.
1 + 2 |E|
wE (E D) = |E| + wpol =
2
2
2

(1.58)
For low frequencies ( << pe ), for which an externally imposed electric
field is shielded out of the plasma, the polarization energy density is dominant.
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

25

In contrast, for high frequencies ( >> pe ) the electron inertia effects cause
the polarization to be small; then, the energy density is predominantly just that
residing in the electric field perturbation itself. The fact that the energy density
caused by electric field perturbations can have a significant (or even dominant,
as occurs for << pe ) component due to the polarizability of the plasma is a
very important aspect of plasma oscillations.

1.6

Ion Acoustic Waves

In the preceding sections we have implicitly assumed that the electrons and
ions both exhibit either adiabatic or inertial responses. However, because the
ions are much heavier, they have a much lower inertial or plasma frequency
and, for the typical case where Te Ti , a much lower thermal speed than
electrons. Thus, for a given length scale x there is an intermediate frequency
regime vT i /x << << vT e /x in which the ions respond inertially while the
electrons respond adiabatically. We will now determine the equation governing
electric field perturbations in a plasma in this regime.
The perturbed electron density for an adiabatic ( << vT e /x) response
induced
by a potential perturbation of a quasineutral plasma equilibrium
=
( s n0s qs = 0) is obtained from the perturbed Boltzmann relation (1.4):
e
.
Te

n
e = n0e

(1.59)

The perturbed ion density for an inertial ( >> vT i /x) response induced by
is obtained from the ion polarization part of the
an electric field perturbation E
total plasma charge density given in (1.41):
#0 2
n
i = pi
qi

dt

t#

$$ ).
dt$$ E(t

(1.60)

The overall perturbed charge density in this intermediate frequency regime is


thus given by
q
#0

!n
s qs

2
+ pi

dt$

#0
0

2
pi

t#

dt

t#

2
$$ ) n0e e
dt$$ E(t
#0 Te

$$ )
dt$$ 2 (t

2De

(1.61)

=
in which we have specialized to electrostatic perturbations for which E
2

and E = .
Substituting this perturbed charge density into Poissons equation (1.1), we
obtain
$
#
0 t
0 t#
1
2
$$ ).
= pi
dt$
dt$$ 2 (t
2 2
De
0
0
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

26

Or, taking the second partial derivative with respect to time, this yields
$ 2
#
1

2
2

= pi
2 .
(1.62)
2
De t2
Considering perturbations
whose
, scale lengths are long compared to the electron
+
Debye length 2 << 1/2De , the equation governing potential perturbations
in the intermediate frequency regime becomes simply
2
c2S 2 = 0,
t2

ion acoustic wave equation,

(1.63)

in which
2
2De =
c2S = pi

Te ni qi2
Zi Te
=
.
mi ne qe2
mi

(1.64)

As indicated in the last equality, for a plasma with a single ion component
ni qi2 = Zi ne e2 so that c2S = Zi Te /mi . The quantity cS has the units of a speed
and as we will see below is the speed of ion acoustic (or sound) waves in a
plasma. It is given numerically by
cS

&

Zi Te
$ 104
mi

'

Zi Te (eV)
m/s,
Ai

ion acoustic speed,

(1.65)

in which Ai is the atomic mass of the ions in the plasma: Ai mi /mp . Here,
we have used the subscript S on c to indicate that these ion acoustic waves
are the natural sound (S) waves that occur in a plasma. The relation of ion
acoustic waves to normal sound waves in a neutral gas are discussed at the last
of this section, and their relation to the sound waves in a magnetohydrodynamic
description of a plasma is discussed in Section 7.2.
Equation (1.63) is a wave equation. In one dimension, say the x direction,
general solutions of it are given by a linear combination of arbitrary functions
f1 , f2 of its mathematical characteristics x cS t:
t) = C1 f1 (x cS t) + C2 f2 (x + cS t),
(x,

where C1 and C2 are arbitrary constants to be fixed by the boundary conditions.


A point of constant phase in this solution moves at the phase speed V of the
wave: d = 0 = dx cS dt = V = dx/dt = cS along the mathematical
characteristics x = x0 cS t.
For wave-like equations such as those in (1.62) or (1.63) we usually seek
solutions of the form
(1.66)
(x, t) = ei(kxt)
in which is a constant, k is the (vector) wavenumber and is the frequency
of the wave. Substituting this Ansatz (proposed form) into (1.62), we find
,+
,
5
4 + 2
2 2
k = 0.
k 1/2De 2 + pi
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

27

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

wpi

k
Figure 1.8: Dispersion diagram for ion acoustic waves in an electron-proton
plasma with Te >> Ti . For kDi << 1 the ion acoustic waves propagate at
the ion acoustic speed: /k $ cS . The dispersion curve = (k) is shown as
> 1 because in this region the ion response is no longer
a dashed line for kDi
inertial (kinetic effects become important) and the present analysis becomes
invalid.

For nontrivial solutions with ,= 0, we must have


2 =

2
pi
k 2 c2S
,
=
2
1 + k 2 De
1 + 1/ (k 2 2De )

ion acoustic wave dispersion relation.


(1.67)

This is called a dispersion relation because it givess the dependence of on k


here for electrostatic ion acoustic waves propagating in a plasma.
The dispersion diagram ( versus k) for ion acoustic waves is shown in
Fig. 1.8. For k 2 2De << 1 (long scale lengths compared to the electron Debye
length) we have /k = cS the phase speed V /k of the wave is the
ion acoustic speed cS . Since we have assumed that the ions have an inertial
response, taking x 1/k we must have vT i /x kvT i << $ kcs . This
condition is satisfied and ion acoustic
1 waves exist in an electron-proton plasma
only if the ion1acoustic speed cS Te /mi is much larger than the ion thermal
speed vT i = 2Ti /mi , which occurs only if Te >> 2Ti . As can be discerned
from (1.67), the wave frequency increases for increasing kDe and asymptotes
to pi for kDe >> 1. However, to satisfy the required condition for an ion
inertial response we must have kvT i << pi or kDi << 1. We can satisfy
kDe >> 1 >> kDi only if Te >> Ti , which is the same as the condition noted
previously in this paragraph for the existence of ion acoustic waves.
As we discussed in the preceeding section, plasma responses can also be
described terms of the plasma giving a dielectric response #. For waves of the
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

28

form given in (1.66) the polarization corresponding to the perturbed charge


density in (1.61) becomes
;
<
2
pi
1

= #0
+ 2 2
E,
(1.68)
P
2
k De
Using the defini = ik and E
= 2 = k 2 .
in which we have used E

tions for the interrelationships between P,


E and # given in (1.53), (1.54), we
find that in the intermediate frequency regime we are considering the plasma
dielectric response is given by
#S (k, ) = #0

2
pi
1
1 2 + 2 2

k De

<

ion acoustic dielectric.

(1.69)

Setting this #S to zero to obtain the normal modes of the plasma readily
yields the dispersion relation for ion acoustic waves given in (1.67). This dielectric function diverges for either 0 or k 0. Thus, again, the plasma
dielectric is only a meaningful quantity for temporal and spatially varying perturbations, i.e., not for an infinite, homogeneous equilibrium. Also, since #S
is real for all real k, (i.e.,the electron and ion components of the polarization are in phase or 180 out of phase with the electric field perturbation), this
intermediate frequency response is also totally reactive (i.e., not dissipative).
Ion acoustic waves are similar to but somewhat different from ordinary sound
,= 0
waves in a neutral gas. Ordinary sound waves are compressible (V

where V is the perturbed flow velocity) mass density perturbations induced by


momentum perturbations propagated by the collisionally coupled flow of the
neutral gas molecules or atoms in response to pressure perturbations see
Section1A.6. They propagate
at a hydrodynamic (H) phase speed given by
1
pn /m = Tn /mn in which = (N + 2)/N is the ratio of the
cH
S =
specific heats, N is the number of degrees of freedom, and pn , m , Tn and Mn
are the neutral gas pressure, mass density, temperature and mass, respectively.
In an electron-proton plasma with Te >> Ti , ion acoustic waves propagate via
,= 0) electric field perturbations, which as we will see in
longitudinal ( E
,= 0, in which
Section 7.2 also lead to compressible flow perturbations V
the adiabatic electron polarization charge density is balanced by an inertial
ion polarization
charge density. Ion acoustic waves propagate at a phase speed
1
cS = Te /mi with the electron temperature coming from the adiabatic electron
Debye shielding and the ion mass coming from the ion inertia. Thus, the physical
mechanism responsible for ion acoustic wave propagation in a plasma is different
from that of sound waves in a neutral gas even though they are both carried by
incompressible flow perturbations collisions couple the atoms or molecules in
a neutral gas whereas the electric field couples electrons and ions together in a
plasma. The ion acoustic speed in a Te >> Ti plasma does not, like ordinary
sound waves, depend on the ratio of specific heats or dimensionality of the
system because it is a one-demensional electric field perturbation rather
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

29

than the collisionally-coupled flow in a neutral gas that propagates ion acoustic
waves in a plasma.

1.7

Electromagnetic Waves in Plasmas

In the preceding three sections we explored the properties of longitudinal (electrostatic) electric field perturbations in an unmagnetized plasma. In this section
we develop the properties of transverse (solenoidal) electric field perturbations
,= 0 but E
= 0 see Sections A.2 and D.5. These types of
for which E
perturbations are often referred to as electromagnetic (em) waves in a plasma
and become light waves in the vacuum limit where the plasma effects are negligible.
To investigate electromagnetic waves in a plasma we begin from the two
Maxwell equations that involve time-derivatives [see (??) in Section A.2]:
#
$
E
,
Amperes law,
(1.70)
B = 0 J + #0
t
B
,
Faradays law.
(1.71)
E =
t
We combine these equations by taking the partial time derivative of Amperes
law and substitute in B/t from Faradays law to obtain
1 2E
J
+ 2 2
t
c t
in which we have used the fact that 0 #0 = 1/c2 , where c is the speed of light
in a vacuum. Since (E) = 2 E ( E), for transverse electric
fields Et ( Et = 0) this can be written as
(E) = 0

1 J
2 Et
.
c2 2 Et =
t2
#0 t

(1.72)

This is a wave equation for the transverse electric field Et . The inhomogeneous
term on the right represents the plasma effects. The general Green function
solution of this equation, including the plasma inertial response effects, is developed in Problem 1.23.
Because electromagnetic waves in a plasma are relatively fast (high frequency) phenomena, we can anticipate that the plasma response will be inertial. Thus, the current perturbation induced by the effect of an electric field
t on the charged particles in a plasma is given by
perturbation E
!
=
s
J
n0s qs v
(1.73)
s

s is the particle velocity perturbation given in (1.32). Taking the


in which v
partial derivative of this current with respect to time, we obtain
!
! n0s q 2

1 J
s
2
2
=
ps
Et =
Et $ pe
Et .
#0 t
m
#
s
0
s
s
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

(1.74)

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

30

[This result can also be obtained by considering the plasma to be a dielectric


medium with the inertial dielectric given by (1.57) and calculating the time
derivative of the displacement current and subtracting off the vacuum contribu t = 2 E

#I /#0 1)E
tion: (1/#0 ) 2 Dt /t2 2 Et /t2 = 2 (
pe t .]
Substituting the resultant inertial plasma response into (1.72), we obtain
t
2E
2
t = 0.
+ pe
Et c2 2 E
t2

(1.75)

This equation is the same as (1.44), which we obtained for electrostatic (or
t
longitudinal electric field) perturbations, except for the presence of the c2 2 E
2
term, which leads to light wave solutions for pe 0. Thus, (1.75) embodies
a combination of charged particle inertial (plasma frequency) and light wave
effects in a plasma.
To explore the properties of electromagnetic waves in a plasma we consider
wave solutions of the form
t ei(kxt) .
t (x, t) = E
E

(1.76)

t into (1.75) yields


Substituting this Ansatz (proposed form) for E
,
+ 2
2
t = 0.
+ k 2 c2 E
+ pe

t ,= 0) solutions are possible for electromagnetic waves that satisfy


Nontrivial (E
2
2 = pe
+ c2 k 2 ,

>
2 / c,
or k = 2 pe

em wave dispersion relation.


(1.77)

This dispersion relation is plotted in Fig. 1.9. Since for these waves /k is
greater than the speed of light and hence, for a nonrelativistic plasma, the thermal speeds of both the ions and electrons, it was valid for us to use the inertial
response that we utilized in (1.74). In the short wavelength limit (k >> c/pe )
the inertial plasma effects become negligible and we have regular light waves
< pe /c), but still high enough frewith $ ck. For longer wavelengths (k
quency so that > pe , the waves have the dispersion characteristics shown
in Fig. 1.9. For pe /c >> k, the waves become electromagnetic plasma oscillations with $ pe . For < pe , the wavenumber k becomes imaginary; this
indicates that transverse electric field perturbations are spatially evanescent in
this regime. In the limit << pe we have k $ i pe /c.
To make the properties of electromagnetic waves in a plasma more concrete,
we consider the propagation of electromagnetic waves from a vacuum into a
plasma. As shown in Fig. 1.10, we consider a situation in which the infinite
half-space where x > 0 is filled with plasma while the infinite half-space where
x < 0 is a vacuum. A wave of frequency is launched from x = in the +x
direction toward the plasma and is incident (I) on the plasma at x = 0. It will

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

31

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

wpe

k
Figure 1.9: Dispersion diagram for electromagnetic waves in a plasma. For
kc << pe the waves become electromagnetic plasma oscillations with $ pe .
For kc >> pe the waves become ordinary light waves with $ ck.
be taken to be of the form:
incident wave:

I eikI xit , E
I = E
t = E
I
ey ,
E
ex , k0 = /c
kI = k0

(1.78)

in which for simplicity we have assumed that the incident wave has linear polarization in the y direction.
In general, part of this wave will be transmitted into the plasma at the
vacuum-plasma interface at x = 0. We take the transmitted (T ) wave to be of
the form
transmitted wave:

T eikT xit ,
t = E
E
ex ,
kT = kT

T = E
T
E
ey ,
>
2 /c
kT = 2 pe

(1.79)

in which the polarization has again been taken to be in the y direction because
the presence of the plasma does not change the wave polarization. In addition,
part of the wave will be reflected; we take the reflected (R) wave to be of the
form
reflected wave:

R eikR xit , E
R = E
t = E
R
ey ,
E
ex , k0 = /c.
kR = k0

(1.80)

The magnetic field accompanying each of these waves is obtained from Fara = ikE
t = B
z =
days law (1.71) for wave solutions of the form (1.76): i B
y / = k E
y /. The boundary conditions at the vacuum-plasma

ez (k
ey ) E
z must be
y and magnetic field B
interface (x = 0) are that the electric field E
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

32

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

Vacuum!Region

Plasma!Region

w!>>!wpe
kI

kT

kR

w!<<!wpe
kI

kR
x!=!0

Figure 1.10: Propagation of an incident (I) electromagnetic wave from a vacuum


into a plasma. For >> pe the wave is transmitted (T) into the plasma with
little reflection (R); the wavenumber k is reduced from /c in the vacuum to
2 1/2
( 2 pe
) /c in the plasma. For << pe the wave is mostly reflected
from the plasma; the part that does penetrate into the plasma is exponentially
evanescent in the electromagnetic skin depth distance c/pe
.

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

33

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA


continuous there. They lead to the two conditions
R = E
T ,
I + E
E
9
:
I E
R = (kT /) E
T .
(k0 /) E

Solving these equations for the relative magnitudes of the transmitted and reflected waves, we find
T
2k0
2
E
>
=
=
,

k0 + kT
2
EI
+ 2 pe
>
2
2 pe
R
k0 kT
E
>
=
=
.
I
k0 + kT
E
+ 2 2

transmitted:

reflected:

pe

The properties of the transmitted and reflected electromagnetic waves are


shown in Fig. 1.10 for two extreme limits: >> pe and << pe . For very
high frequencies ( >> pe ) the incident electromagnetic is transmitted into the
plasma with very little reflection and only a slight reduction in the wavenumber
k. As the>
frequency of the incident wave is decreased, the wavenumber decreases
2 /c. At the point where = , the transmitted wave has
to kT 2 pe
pe
oscillation. For < pe
kT = 0 and becomes just an electromagnetic plasma
>
2 2 /c. The plus sign is
the wavenumber becomes imaginary, kT = i pe

the physically relevant solution since it leads to evanescence (spatial decay not
due to a dissipative process) in space for x > 0. In the limit << pe the
incident wave is mostly reflected and the small component of the wave that is
transmitted into the plasma is given by
T exp[x/(c/pe ) it],
t $ E
E

<< pe .

(1.81)

This electric field perturbation is exponentially evanescent in the distance e


given by
e c/pe ,

electromagnetic skin depth.

(1.82)

Thus, for > pe electromagnetic waves are partially reflected at the


vacuum-plasma interface and propagate into plasmas with some reduction in
the wavenumber k. However, for < pe the plasma (and in particular the
electron) inertial response to an electromagnetic wave causes the wave to be
mostly reflected at the vacuum-plasma interface and prevents
> the wave from
2 2 , which
penetrating into a plasma more than a distance of about c/ pe

becomes just the electromagnetic skin depth c/pe in the limit << pe .
A major diagnostic application of the properties of electromagnetic waves in
a plasma is their use in a microwave interferometer to determine the density of
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

34

reference!leg
attenuator
microwave
source

phase
detector

plasma

x
0

Figure 1.11: Schematic illustration of a microwave interferometer. The electromagnetic wave passing through the plasma has a smaller wavenumber (longer
wavelength) than the wave passing through vacuum in the reference leg. Thus,
there is a phase shift between the two signals arriving at the detector.

a plasma see Fig. 1.11. The difference in the phase between the wave that
passes through a reference vacuum leg versus the wave that passes through a
leg with a plasma in it is given by
>

0 L
0 L
2 (x)
2 pe
.
dx [kI kT (x)] =
dx
=
c
0
0

In the limit >> pe this becomes simply


$

2
(x)
pe
e2
=
dx
2c
2me #0 c

dx ne (x).

(1.83)

Since the square of the electron plasma frequency is proportional to the local
plasma density, the measurement of this phase shift determines the line integral
of the electron density in the plasma. For example, microwave interferometers
with frequencies in the 50 200 GHz range are commonly used to measure the
(L
line-average density n
e (1/L) 0 dx ne (x) of plasmas with electron densities
in the 1018 1020 m3 range. For some other applications in which the properties
of electromagnetic waves in plasmas are important see Problems 1.271.29.

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

1.8

35

Plasma Definition and Responses

Now that we have elucidated the basic length (Debye length D ) and time
(plasma period 1/pe ) scales for collective phenomena in plasmas, we can specify
quantitatively the criteria that must be satisfied for matter to exist in the plasma
state. As we discussed in the introduction to this chapter, a general criterion for
the existence of a plasma is that charged particle interactions be predominantly
collective rather than binary in the medium. For this general criterion to be
satisfied, we must require that it be satisfied in charged-particle interactions, as
well on the relevant length and time scales for collective phenomena:
1. n3D >> 1. The number of charged particles within a Debye cube (or
sphere) must be large so that: a) collective interactions dominate over
binary interactions at the mean interparticle separation distance; b) the
energy density embodied in the polarization electric field around a given
charged particle is small compared to a typical particles kinetic energy;
and c) the thermal noise level is small see (1.11), (1.14) and (1.15),
respectively.
2. L >> D . The spatial extent of a collection of charged particles must
be large compared to the collective interaction scale length for plasmas,
the Debye length D , so that: a) the collective interactions are dominated
by bulk plasma rather than boundary effects; and b) inertial effects are
determined locally see (1.27) and Fig. 1.3, and (1.49), respectively.
3. pe >> en . The collective inertial response frequency in a plasma, the
electron plasma frequency pe , must be large compared to the electronneutral collision frequency en , so that the fundamental inertial responses,
the electrostatic electron plasma oscillations in (1.45) and the plasma oscillation effects on electromagnetic waves in (1.75), are not damped by
dissipative neutral particle collision effects.
While we have derived the basic collective phenomena in an unmagnetized
plasma, the same physical phenomena occur in magnetized plasmas (primarily along the magnetic field direction); hence these criteria for the existence of
the plasma state apply to magnetized plasmas as well.
Among the three criteria for existence of the plasma state, the first one,
the requirement that there are many charged particles in a Debye cube, is the
necessary condition and the most critical. After this fundamental criterion is
satisfied, the second and third criteria are just checks (sufficient conditions)
that the behavior of the medium will be dominated by collective plasma phenomena on the basic plasma length and time scales. The fundamental plasma
parameter, the number of charged particles in a Debye cube, depends on the
plasma temperature and charged-particle density, i.e., n3D T 3/2 /n1/2 . Thus,
as shown in Fig. 1.12, we can exhibit the various types of plasmas that occur in
nature by showing where they lie relative to lines of constant n3D in a plot of
electron temperature versus electron plasma density. As shown in this figure,
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

36

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

Plasma!Frequency!fpe!(Hz)
10

10

10

10

101!m

10

14

10-5!m

16

solar
corona

10

inertial
fusion

10

10

10

solar!atmosphere
MHD
generators

flames

glow
discharges

10

10

15

partially
ionized

alkali!metal
plasmas

ionosphere

interplanetary
interstellar

10-2

10

10-8!m

10

18

10

10-11!m

magnetic!confinement
experiments

10

earths
magnetosphere

10

Temperature!(eV)

12

10

10-2!m

104

10

10

magnetic
fusion

=!104!m

inertial
confinement
experiments

10

10

10

20

10

25

30

10

10

35

Electron!Density!(m-3)
Figure 1.12: Ranges of electron temperature and density for various types of laboratory and extraterrestial plasmas. Also shown are the characteristic plasma
parameters: electron Debye length De (constant along the dashed lines), number of charged particles in a Debye cube ne 3De (constant along solid lines), and
electron plasma frequency pe (constant along vertical lines). Also indicated is
the electron temperature range below which the medium is not fully ionized,
which is determined from the Saha equation [see Section A.7 and in particular
(??) and (??)].

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

37

the plasma state spans an enormous parameter range 26 orders of magnitude


in density and 7 orders of magnitude in temperature!
=
Almost all plasmas are electrically quasineutral (i.e., q = s ns qs $ 0) on
length scales longer than the Debye length D . (Notable exceptions are the
electron-rich, non-neutral magnetized plasmas see references listed below.)
On length scales shorter than a Debye length the local charge density and potentials are dominated by the effects of the discrete charged particles. However,
on length scales longer than the Debye length the collective, plasma polarization effects dominate and the plasma is quasineutral. The use of a quasineutral
approximation for scale lengths longer than a Debye length is often called the
plasma approximation.
Slow processes ( << vT /x kvT ) in a plasma are governed by the adiabatic response, as discussed in Section 1.1. Fast processes ( >> vT /x kvT )
are governed by the inertial response, as discussed in Section 1.4. Because
the electron
1 thermal speed is usually much greater than the ion thermal speed
(vT e mi /me vT i 43 vT i >> vT i for Te Ti ), the electrons and ions in
a plasma can respond differently to perturbations for kvT i << << kvT e
electrons respond adiabatically while the ions respond inertially, as discussed in
Section 1.4.
The response of the plasma to electric field perturbations leads to polarization of the plasma, and hence to a dielectric response for the plasma medium.
The plasma responses in the various frequency regimes can be summarized in
terms of the density and dielectric responses to small (i.e., linearizable) wavelike
perturbations of the form exp (ik x it) in an infinite, homogeneous electronion plasma as follows:
Adiabatic (A) electrons and ions: / kvT i , kvT e ; Debye shielding;
"
#
$%
q
1
1
1
+ 2
n
A $ n0 ; #A (k, ) $ #0 1 + 2
.
T
k
2De
Di

(1.84)

Adiabatic electrons, inertial ions: kvT i / / kvT e ; ion acoustic waves;


.
/
2
2
pi
pi
e
Zi e
niI
1

n0e ,
$ 2 E; #S (k, ) $ #0 1 + 2 2 2 .
n
eA $
Te
#0

k De

(1.85)

Inertial (I) electrons and ions: kvT i , kvT e / ; plasma oscillations;


.
/
2
2
2
pi
ps
pe
sI
qs n

$ 2 E;
#I (k, ) $ 1 2 2 .
#0

(1.86)

As can be seen from these various responses, a plasma is an electrically active medium with a frequency- and wavenumber-dependent polarizability and
dielectric response function. As discussed before, these responses are only applicable for spatially and temporally varying perturbations they all diverge
for , k 0.
DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

38

Within the approximations employed in this chapter, all the basic phenomena in plasmas that we have discussed are reactive with no dissipation. Dissipation would be caused by polarization components that are 90 out of phase
with the electric field perturbations, which for exp (ik x it) perturbations
would be indicated by an imaginary part of the dielectric #.
Implicitly, we have been considering the plasma to be collisionless. Presuming collisions with neutrals are negligible (p >> n ), there are two types
of effects that lead to evanescence of waves in a plasma Coulomb collisions,
which will be discussed in Chapter 2, and wave-particle resonance effects (Landau damping), which will be discussed in Chapter 8. Since the thermal noise
fluctuation energy induced by two-particle
(or Coulomb collisions)
,
+ correlations
in a plasma is only a small fraction 1/ n3D << 1 of the thermal energy in
a plasma, we can anticipate
that
the average Coulomb collision frequency will
+
,
also be small: p / n3D << p . Cumulative small-angle Coulomb
+
, collisions enhance the Coulomb collision rate by a factor of order ln n3D see
Chapter 2 but do not change the basic conclusion that the Coulomb collision
rate is slow in a plasma as long as n3D >> 1. The wave-particle resonance
the wave phase speed /k is
effects will be largest when kvT , i.e., when 1
of order the most probable thermal speed vT 2T /m of one of the species
of charged particles in a plasma. Thus, wave-particle resonance effects will lead
to evanescence of waves (Landau damping) for /k vT i or vT e . These wave
phase speeds and corresponding frequencies are between the frequency ranges
we have considered in this chapter and require a kinetic plasma description.
Wave-particle resonance effects and Landau damping are discussed in Chapter
8, and in particular in Section 8.2.
REFERENCES AND SUGGESTED READING
Discussions of plasma sheath and Langmuir probe theory can be found in
D. Bohm in Characteristics of Electrical Discharges in Magnetic Fields, A.
Guthrie and R.K. Wakerling, eds. (1949) [?].
F.F. Chen, Electrical Probes, in Plasma Diagnostic Techniques, R.H. Huddlestone and S.L. Leonard, eds. (1965), Chapt. 4 [?].
L. Schott and R.L.F. Boyd in Plasma Diagnostics, W. Lochte-Holtgreven, ed.
(1968) [?].
J.D. Swift and M.J.R. Schwar, Electric Probes for Plasma Diagnostics (1971)
[?].
P.M. Chung, L. Talbot and K.J. Touryan, Electric Probes in Stationary and
Flowing Plasmas (1975) [?].
P.C. Stangeby, The Plasma Sheath, in Physics of Plasma-Wall Interactions
in Controlled Fusion, D.E. Post and R. Behrisch, eds. (1985), Vol. 131, p. 41
[?].
Hutchinson, Principles of Plasma Diagnostics (1987), Chapt. 3 [?].
N. Hershkowitz, How Langmuir Probes Work, in Plasma Diagnostics, O. Auciello and D.L. Flamm, eds., (1990) [?].

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

39

Recent books that discuss the various types of plasmas indicated in Fig. 1.12,
some of which are beyond the scope of this book, include
Magnetically confined plasmas for controlled fusion:
Rose and Clark, Jr., Plasmas and Controlled Fusion (1961) [?].
Miyamoto, Plasma Physics for Nuclear Fusion (1980) [?].
Miyamoto, Fundamentals of Plasma Physics and Controlled Fusion (1997) [?].
Stacey, Fusion Plasma Analysis (1981) [?].
Teller, ed., Fusion (1981), Vol. I, Parts A & B [?].
Nishikawa and Wakatani, Plasma Physics, Basic Theory with Fusion Applications (1990) [?].
Hazeltine and Meiss, Plasma Confinement (1992) [?].
White, Theory of Tokamak Plasmas (1989) [?].
Goldston and Rutherford, Introduction to Plasma Physics (1995) [?].

Laser-produced plasmas for inertial confinement fusion:


Kruer, The Physics of Laser Plasma Interactions (1988) [?].
Lindl, Inertial Confinement Fusion (1995) [?].

Space plasmas:
Parks, Physics of Space Plasmas, An Introduction (1991) [?].
Gombosi, Physics of the Space Environment (1998) [?].

Cosmic plasmas:
Alfven and F
althammer, Cosmical Electrodynamics, Fundamental Principles
(1963) [?].
Parker, Cosmical Magnetic Fields, Their Origin And Their Activity (1979) [?].
Sturrock, Plasma Physics, An introduction to the theory of astrophysical, geophysical, and laboratory plasmas (1994) [?].
Choudhuri, The Physics of Fluids and Plasmas, An Introduction for Astrophysicists (1998) [?].

Partially ionized plasmas and plasma processing:


Chapman, Glow Discharge Processes (1980) [?].
Lieberman and Lichtenberg, Principles of Plasma Discharges and Materials Processing (1994) [?].

Nonneutral plasmas:
Davidson, Physics of Nonneutral Plasmas (1990) [?].
Marshall, Free Electron Lasers (1985) [?].

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

40

PROBLEMS
1/3

1.1 Evaluate the scale lengths bcl


and De for an electron-proton plasma
min , ne
in a typical small-scale magnetic confinement experiment (e.g., a universitybased tokamak) with ne = 2 1019 m3 , Te = Ti = 300 eV. Compare them
to the Bohr radius, de Broglie wavelength, and the classical electron radius
re = e2 /({4#0 }me c2 ). Discuss the physical significance of each of these scale
lengths. Over what length scale will collective effects occur in such a plasma? /
1.2 Calculate the plasma parameter n3D for the plasma described in the preceding
problem. Estimate the effective temperature for thermal noise in such a plasma.
Compare this thermal noise temperature to normal room temperature. /
1.3 Consider a hypothetical situation in which all the electrons in a homogeneous
and quasineutral but bounded plasma are displaced a small distance x in the
ex
direction. Show that in the bulk of the plasma the electric field is unchanged,
but that in a layer of width x at the plasma edge there is an electric field.
How large a displacement x induces a maximum potential change equal to the
electron temperature in the plasma? Compare this length to the electron Debye
length De and discuss why such a comparison is relevant. //
1.4 Determine the one-dimensional potential distribution in a plasma around an
infinite sheet charge with a one-dimensional surface test charge density given
by q = t (x xt ). ///
1.5 Show that for a two-dimensional situation of an appropriately modified form of
(1.7) the potential around a line charge in a plasma is given by
t (x) = (2t /{4#0 })K0 (|x xt |/D )
in which t is the line charge density (coulombs/m) for a line charge of infinite
length placed at x = xt and K0 is the modified Bessel function of the second
kind of order zero. ///
1.6 Show that the potential given by (1.10) is the Green function for the adiabatic
(Debye shielding) response to a free charge density free (x) in an infinite, homogeneous plasma, and thus that the general potential solution is given by
0
free (x" ) exp(|x x" |/D )
(x) = d3x"
.
{4#0 } |x x" |
Discuss the physical scale lengths over which this Greens function solution is
valid. Compare this result to the corresponding potential induced by a charge
density in vacuum given in (??). ///
1.7 Show that the combination of the charge of a test particle and the polarization
charge density it induces produces a vanishing net charge Q in the plasma. //
1.8 A spherical spacecraft orbiting the earth in a geostationary orbit finds itself
immersed in a plasma that typically has an electron density of about 106 m3
and temperature of about 100 eV. Sketch the spatial variation of the electric
potential around the spacecraft, indicating the magnitudes of the potential and
spatial scale lengths involved. To what potential does the spacecraft charge up
relative to its surroundings? /

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

41

1.9 A spherical probe 3 mm in diameter is inserted into a fully ionized electronproton plasma generated by microwave heating power applied to a hydrogen gas
and has ne = 1015 m3 , Te = 10 eV, Ti = 1 eV. What is the Debye length in
such a plasma? Is it large or small compared to the probe size? If the probe is
biased to 10 V, how much current would it draw? /*

1.10 It is proposed to put a wire screen into the plasma described in the preceding
problem and bias it so as to exclude plasma from the region behind the screen.
Taking account of sheath effects, how closely spaced must the screen wires be
so as to block a substantial fraction of the plasma? To what potential should
the probe be biased? /*

1.11 The electron temperature in low energy density plasmas can be measured with a
double probe a single assembly with two identical but differentially biased
Langmuir probes that electrically floats and draws no net current from the
plasma. Show that the current flowing between the two probes is given by
I = 2 ISi tanh (e/2Te )
in which is the potential difference (voltage) between the two probes. //*
1.12 The static electrical admittance Y (inverse of impedance Z) of a sheath is
given by I/B . At what bias potential B should this partial derivative
be evaluated? Show that the sheath admittance is given approximately by
Y # I/(Te /e). Up to what frequency will this estimate be valid? //*

1.13 For large negative wall potentials (|W | >> Te /e) applied between two grids
in a planar diode the electrical current is limited by space charge effects. Derive the Child-Langmuir law for this limiting current for a grid separation d as
follows. First, using (1.27) show that when the sheath thickness xS d and
W , the ion speed V that corresponds to the space-charge-limited ion
flow at the sheath edge can be written in terms of the wall potential W and
the grid separation d. Then, show that the (ion) current density into the sheath
region between the grids is given by
&
# $# 2 $
4 2e #0 |W |3/2
4
D
J=
= (ne eVW )
9 mi
d2
9
d2
in which VW (2e|W |/mi )1/2 is the ion speed at the wall, and D =
(#0 |W |/ne e)1/2 is an effective Debye length. //*

1.14 In the Plasma Source Ion Implantation (PSII) technique [J.R. Conrad, J.L.
Radtke, R.A. Dodd, F.J. Worzola, N.C. Tran, J. Appl. Phys. 62, 4591 (1987)],
the target to be bombarded is inserted into a plasma with parameters ne
109 cm3 and Te 2 eV, and a natural sheath is allowed to form around it.
Then, the target is rapidly biased to a very large ( 30 kV) negative potential
B . This expels the lighter electrons from the region around the object, which
in turn causes an ion matrix to be formed there. On what time scales are
the electrons expelled, and the new sheath formed? What is the approximate
maximum energy and current density of the ions bombarding the target before
the new sheath forms? Compare this current density to that given by the ChildLangmuir law discussed in the preceding problem. Finally, estimate the fluence
(ions/cm2 ) per pulse and the number of pulses required to inject an atomic
monolayer of ions in the target. ///*

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

42

1.15 Consider an impure laboratory plasma composed of a number of different types


of ions: protons with np = 3 1019 m3 , fully ionized carbon ions with 10% of
the proton density and iron ions that are 23 times ionized (Lithium-like charge
states) with 1% of the proton density. What is the electron density and the
overall plasma frequency in this plasma? Also, what is the dielectric constant
for 90 GHz electrostatic fluctuations in this plasma? /
1.16 Consider electrostatic plasma oscillations in an electron-positron plasma, such
as could occur in interstellar space, with ne = ne+ = 106 m3 . What is the
plasma frequency for such oscillations? Assume the electrons and positrons have
temperatures of 100 eV and that the cross-section for an annihilation interaction
between them is given by times the square of the classical electron radius
re = e2 /({4#0 }me c2 ), i.e., an re2 . What is their annihilation reaction
rate? Compare this annihilation rate to the plasma frequency. /
1.17 Consider a situation where an oscillating potential is applied across two plates
on either side of a plasma such as that described in Problem 1.9. Assume the
plates are separated by 10 cm and that the potential oscillates at 100 kHz.
What is the dielectric constant for these oscillations? What is the ratio of
the energy density in the plasma polarization fluctuations to that in the electric
field fluctuations? /
1.18 Calculate the weak dissipation induced by Coulomb collisions ( << pe ) of
pe electrostatic oscillations in a plasma as follows. Add a collisional dynamical
=E
sin t the
friction force me v [cf., (??)] to (1.31) and show that for E
perturbed velocity of electrons is then given for t >> 1/ by
v#

eE
[cos t (/) sin t].
me

Next, calculate the average of the Joule heating in the plasma by the oscillations
E)
t . Finally, use a wave energy balance
over an oscillation period 2/, i.e., (J
equation [cf., (??)] to show that
E)
t
(wE )t
(J
1
2
=
#
2
(wE )t
t
(wE )t
1 + 2 /pe
in which (wE )t is the average of the electrostatic wave energy density in the
plasma over an oscillation period. ///
1.19 Consider a hypothetical situation in which all the electrons in a thin slab are
displaced a small distance x0 in the
ex direction. Show that the electric field
x in the region where
induced by this displacement is given by E = (n0 e/#0 ) x e
the electrons are displaced. Then, show from Newtons second law that this
force causes the position of the slab of displaced electrons to oscillate at the
electron plasma frequency. //
1.20 Taking account of plasma sheath effects, sketch the spatial variation of the
potential (x) between the plates of a capacitor filled with plasma assuming
the capacitor has a potential 3 Te /e applied across it. Next, consider a
case where an oscillating potential = 0 sin t is applied across the plasma
capacitor with 0 = 10 Te /e and = pe /10. If the capacitor plate separation
is L (>> D ), how large is the electric field component oscillating at frequency
in the body of the plasma? //

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

43

1.21 Show that for a one-dimensional wave perturbation in a plasma with E(x,
t) =

E
ex sin(kx t) the nonlinear terms in (1.31) are negligible in (1.32) when the

wave-induced velocity jitter in the particle motion, vjitter q E/m,


is small
compared to the wave phase speed /k, or alternatively when k
xjitter << 1. //
1.22 Consider the propagation of ion acoustic waves in a typical hollow cathode arc
discharge composed of electrons and doubly charged Argon ions with ne = 1019
m3 , Te = 10 eV, Ti = 1 eV. Discuss why the conditions for propagation of
ion acoustic waves are satisfied in this plasma. What is the ion acoustic speed
in this plasma? Compare it to the speed of sound in air at the earths surface.
With what wavelength and phase speed will externally imposed waves with a
frequency of 100 MHz propagate in this plasma? /
t induced by a small free current Jfree
1.23 Show that the transverse electric field E
in a plasma is governed by the equation
#
2 $
2
pe
t 1 Et = 0 Jfree .
2 2 E
2
c
c t2
t
Then, show that a Greens function solution of this equation in an infinite,
homogeneous plasma which satisfies this equation is
"
#
$%
0
Jfree (x" , t" )/t"
|x x" |
t = 0
E
d3x"
exp

4
|x x" |
c/pe
ret
in which the square bracket [ ]ret means that the time t" is to be evaluated at
the retarded time t" = t |x x" |/c. ///

1.24 Use the solution in the preceding problem to calculate the transverse electric
t caused by the current qt v[x x(t)] produced by a nonrelativistic test
field E
particle moving along the trajectory x = x(t) in a plasma. Show that this
transverse electric field points in the direction of test particle motion. Also,
show that for |x x(t)| < D its magnitude is of order v 2 /c2 << 1 compared
to the longitudinal electric field produced by the electrostatic potential t in
(1.10). ///
1.25 Plot the wave dispersion diagrams for electrostatic ion acoustic waves and electromagnetic plasma waves (i.e., Figs. 1.8 and 1.9) in the plasma described in
Problem 1.22 on a single versus k diagram with approximately linear scales.
Indicate in which regions of this diagram adiabatic and inertial responses for
the electrons and ions are applicable. /
1.26 A 140 GHz microwave interferometer set up across a 30 cm thick column of
plasma measures a phase shift of 240 . What is the line-average plasma
density in the column? /
1.27 Amateur radio operators routinely communicate via shortwave radio over long
distances around the earth. Since communication by direct line of sight is not
possible because of the curvature of the earths surface, the waves must be reflected from the ionosphere above the earths surface. What frequency range
corresponds to the 10 to 40 meter free space wavelength range used by amateur radio operators for these communications? What is the minimum electron
density and height of the ionosphere above the earths surface for single-bounce
communications over the approximately 6000 km from the United States to
Western Europe? /

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 1. COLLECTIVE PLASMA PHENOMENA

44

1.28 During reentry of satellites into the earths upper atmosphere, microwave communications in the 300 MHz frequency range are blacked out by the plasma
formed in the heated air around the satellite. How high must the plasma density be around the satellite and how thick must the plasma be to cause the
communications blackout? /
1.29 In one type of inertial fusion experiment, intense light from a laser is shined on
a frozen hydrogen pellet. As the laser light is absorbed it heats up the pellet
and produces a plasma on its surface. Light from a Neodynium glass laser ( =
1.06 m) is ultimately observed to be reflected from the pellet. How high must
the density of free electrons be in the plasma around the pellet? Compare
this density to the original solid density of the pellet. How thick must the layer
of free electrons be to reflect (or refract) the light waves? Compare this length
to a typical pellet radius of 3 mm. /
1.30 In plasma processing of materials for the semiconductor industry an inert, low
pressure gas is partially ionized by radiofrequency waves in a vacuum chamber.
Consider a case where the initial gas is Argon at a pressure of 104 mm Hg (a
760 mm column of mercury corresponds to atmospheric pressure), the electron
density is 107 cm3 , the electron temperature is 3 eV and the temperature of the
singly charged Argon ions is 0.1 eV. What is the degree of ionization in this gas?
Estimate the electron-neutral collision frequency en using an electron-neutral
cross-section of 103 a20 where a0 is the Bohr radius. Does this medium satisfy
all the criteria for being a plasma? How large must it be to satisfy the length
criterion? /
1.31 An oscillating potential of 3 volts at a frequency of 1 MHz is applied to a probe
inserted into the plasma described in Problem 1.9. Over what distance ranges
from the probe can adiabatic or inertial responses be used for the electrons and
for the ions in this plasma? /
1.32 What is the dielectric constant for externally imposed waves with a frequency
of 1 MHz and a wavelength of 5 cm in the plasma described in problem 1.9?
What would the dielectric constant be if the wavelength was increased to 500
cm? /

DRAFT 10:26
August 12, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

Chapter 2

Coulomb Collisions
The characteristics and effects of Coulomb collisions between charged particles
in a plasma are very different from those of the more commonly understood
collisions of neutral particles. The fundamental differences can be illustrated by
examining trajectories of neutral and charged particles as they move through a
partially ionized gas. As shown in Fig. 2.1, neutral particles move independently
along straight-line trajectories between distinct collision events. Collisions occur
when neutral atoms or molecules come within an atomic radius (of order 1
A=
1010 m see Section A.7) of another particle (a neutral or a charged particle)
and the electric field force associated with the atomic potential (of order eV) is
operative; the resultant strong, typically inelastic, collision causes the initial
neutral to be scattered in an approximately random direction.
In contrast, as a charged test particle moves through an ionized gas it simultaneously experiences the weak Coulomb electric field forces surrounding all
the nearby charged particles, and its direction of motion is deflected as it passes
by each of them, with the closest encounters producing the largest deflections
see Fig. 2.2. As was discussed in Section 1.2, the Coulomb potential around
any particular background charged particle in a plasma is collectively shielded
out at distances beyond a Debye length. Thus, the only background particles
that exert a significant force on the test particles motion are those within about
a Debye length. However, because plasmas usually have a very large number
of particles within a Debye sphere [(4/3)n3D >> 1], even in traversing only
a Debye length the test particles motion is influenced by a very large number
of background particles. The Coulomb electric field forces produced by individual background particles are small and can be assumed to be experienced
randomly by the test particle as it passes close to individual background particles as indicated in the electron trajectory shown in Fig. 2.2. The effect of
many successive, elastic Coulomb collisions of a test particle with background
charged particles leads to a random walk (Brownian motion) process. Thus,
the effects of the many cumulative small-angle, elastic Coulomb collisions are
diffusion of the test particles direction of motion (at constant energy in the
center-of-momentum frame) and consequently deceleration of the test particles
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

++

-+ +
o

+o
-

o+ --

+
+

+
+

+- +

-+ + +
-+
o+
+

+
+
+
+
+
oo
+
o +- +
o
+
+
+
- +
+ + --+ o +
+
+
+
+ +
o
+
++
-+
++
+
- +
+
-o
-+o + +
o
o
+
-+
+
o
+o
-

Figure 2.1: The trajectory of a neutral particle in a partially ionized gas exhibits
straight-line motion between abrupt atomic collisions. In this and the next
figure, the (assumed stationary) random positions of background particles in
the partially ionized plasma that the incident test particle could interact with
are indicated as follows: neutral particles (circles), electrons (minus signs) and
ions (plus signs).

++

-+ _ +
o

+o
-

o+ --

+
+

+
+

+- +

-+ + +
-+
o+
+

+
+
+
+
+
oo
+
o +- +
o
+
+
+
- +
+ + --+ o +
+
+
+
+ +
o
+
++
--+
+
+
+
- +
- +
o
-o
- o
- +o + +
+
+
+
o
+o
+

Figure 2.2: The trajectory of a test charged particle (electron) in a partially


ionized gas exhibits continuous small-angle deflections or scatterings of its direction of motion. The largest deflections occur when a charged particle passes
close to another charged particle.

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

initial, directed velocity. Exploration of these Coulomb collision effects is the


main subject of this chapter.
Because electrons have less inertia and typically have larger speeds than
ions, their collision rates are usually the largest in plasmas. Thus, we first consider the momentum loss and velocity-space diffusion of a test electron as it
moves through a plasma. Electron collisions are investigated using the Lorentz
(simplest) collision model in which their collisions are assumed to occur only
with a background of stationary ions. Next, since the collisional effects decrease with electron speed, we determine the energy (usually on the high energy
tail of a Maxwellian distribution) at which electrons run away in response to
an electric field; also, the plasma electrical resistivity is determined by balancing, for the backgound electron distribution, the average collisional deceleration
against the electron acceleration induced by an electric field. Then, we discuss
phenomenologically the various Coulomb collisional processes (momentum loss,
energy exchange and their time scales) that occur between electrons and ions in
a plasma. The chapter concludes with sections that develop a more complete
model of Coulomb collision effects that takes into account collisions with all
types of background charged particles that are also in motion. Finally, applications of this more complete model to the thermalization of a fast ion in a plasma
and to the evolution of the velocity of any type of test particle are discussed.

2.1

Lorentz Collision Model

To illustrate Coulomb collision effects, we first consider the momentum loss and
velocity diffusion of a test electron moving through a randomly distributed background of plasma ions that have charge Zi e and are stationary. (The particles
in the background that are being collided with are sometimes called field particles.) The background plasma electrons, which must be present for quasineutrality, will be neglected except insofar as they provide Debye shielding of the
Coulomb potentials around the background ions. However, the test electron
can be thought of as being just one particular electron in the plasma. This
simplest and most fundamental model of collisional processes in a plasma is
called the Lorentz collision model . It provides a reasonably accurate description of electron-ion collisional processes and, in the limit Zi >> 1 (see Section
2.8) where electron-electron collisional effects become negligible, for electron
Coulomb collision processes as a whole.
The electron test particle velocity v will be assumed to be large compared to
the change v due to any individual Coulomb interaction with an ion. Hence,
the test electron will be only slightly deflected from its straight-line trajectory
during a single collision. Figure 2.3 shows a convenient geometry for describing
the Coulomb collision process.1 In the rest frame of the electron, the background
ion, which we place at the origin of the coordinate system, is seen to be moving
1 The geometry shown in Fig. 2.3 and the pedagogical approach we use for exploring
Coulomb collision processes follows that developed in Chapter 13 of Jackson, Classical Electrodynamics, 1st and 2nd Editions (1962, 1975).

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS


x

electron

b!cos!

v
ion

vt

b!sin!

Figure 2.3: Geometry for considering the Coulomb collision of an electron having
charge qe = e with an ion of charge qi = Zi e. The ion is placed at the origin
of the coordinate system, which in the electron rest frame is moving in the
ez
direction at the electron speed v. The electron passes the ion at a distance b at
the closest point, which occurs at t = 0.

ez . The electron
with a velocity v
ez along a straight-line trajectory x(t) = vt
is instantaneously at the position
!
"1/2
ey sin ) + vt
ez , |x| = b2 + v 2 t2
,
(2.1)
x = b (
ex cos +

in which b is known as the impact parameter . It is the distance of closest


approach, which by assumption will occur at time t = 0. The electrostatic
potential around the ion is the Coulomb potential (x) = Zi e/({4%0 }|x|).
Thus, the electric field force experienced by the test electron with charge qe = e
at its position x is
#
$
Zi e
Zi e2 x
=
.
(2.2)
F = qe E = (e)
{4%0 }|x|
{4%0 }|x|3

Next, we calculate the momentum impulse me v on the test electron as it


passes the background ion. Integrating Newtons second law (m dv/dt = F)
over time from long before (t ) to long after (t +) the Coulomb
collision that takes place during the time t where |t| t b/v, we see that
a single electron-ion Coulomb collision induces:
%
%
Zi e2 x
dt qe E =
dt
.
(2.3)
me v =
{4%0 }|x|3

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS


Using the specification of x in (2.1), we find
v

Zi e2 b
(
ex cos +
ey sin )
{4%0 }me

(b2

dt
+ v 2 t2 )3/2

2Zi e2
(
ex cos +
=
ey sin ) .
{4%0 }me bv

(2.4)

(This expression is&relativistically correct if me is replaced by the relativistic


mass me = me / 1 v 2 /c2 .) Note that the perturbation of the electron velocity is in a direction perpendicular to its direction of motion. There is no
component along the direction of particle motion (
ez direction), at least in this
first order where the particle trajectory is the unperturbed one because the z
component of the Coulomb force is an odd function of z or t. Hence, to this first
or lowest order there is no momentum loss by the particle. Rather, a typical
electron is only deflected by a small angle v /v << 1 in velocity space.
1/3
Using a typical impact parameter b ne
, the average inter-particle spacing
in the plasma, and a typical electron speed v vT e , the typical deflection angle
is 1/[4(ne 3D )2/3 ] << 1.
Since the background ion is at rest in the Lorentz collision model, the electron
energy is conserved during the elastic Coulomb collision process. Thus, we have
me |v|2 /2 = me |v + v|2 /2 = m(|v|2 + 2v v + v v), from which we find
that the component of v parallel to v can be determined from
1
1
v v = v v & v v ,
2
2

(2.5)

as indicated in Fig. 2.4. That is, for electron energy conservation the change of
electron velocity along its direction of motion is given by half of the negative of
the square of the transverse deflection. The net velocity change along the
ez or
parallel (') direction of electron motion which is induced by a single Coulomb
collision with a background ion is thus (v v v v$ )
v$ &

1
2Zi2 e4
v v =
.
2v
{4%0 }2 m2e b2 v 3

(2.6)

Note that while v is a first order quantity in terms of the weak Coulomb
electric field between the two particles given in (2.2), v$ is a second order
quantity, as evidenced by the square of the {4%0 } factor in the denominator.
The result in (2.6) can also be obtained directly by integrating the Coulomb
electric field force along a perturbed (by the Coulomb collision) trajectory rather
than the straight-line electron trajectory that was assumed in the preceding
analysis see Problem 2.5.
Next, we take account of the entire background distribution of ions, assuming
that the electron collisions with individual ions are statistically random and
thus that their effects can be summed independently. For a density ni of ions,
adopting a cylindrical geometry in which the radius is b and the azimuthal
angle
'
is , the number of ions passed by the electron per unit time is ni d3x/dt =
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

constant
energy
surface

v!+!v

v
v"

Figure 2.4: Change in electron velocity vector from v before the Coulomb collision to v + v afterward. The change takes place at constant electron energy,
which means constant radius in this diagram, and hence results in v$ < 0.
'
'
'
ni (dz/dt) dA = ni v d b db (cf., Fig. 2.2). Hence, the net or ensemble
average2 Coulomb collisional force in the direction of electron motion is given
by
% 2 %
%
)v$ *
4 ni Zi2 e4
db
= ni v
d
b db me v$ =
. (2.7)
)F$ * me
2 m v2
t
{4%
}
b
0
e
0
0
Here, t is a typical interaction time for individual Coloumb collisions (t
b/v 1/[pe (ne 3D )1/3 ]), which is short compared to the time for the test electron to traverse a Debye sphere ( De /v De /vT e 1/pe ), and certainly
short compared to the time scale on which the test particle velocity v changes
significantly due to Coulomb collisions [t << 1/, where is the collision
frequency defined in (2.14) below].
The integral over the' impact parameter b in (2.7) is divergent at both its

upper and lower limits: 0 d b/b = ln(/0) ?! We restrict its range of integration through physical considerations that can be more rigorously justified by
detailed analyses. The maximum impact parameter will be taken to be the Debye length since the Coulomb electric field force decays exponentially in space
2 In an ensemble average one averages over an infinite number of similar plasmas (realizations) that have the same number of particles and macroscopic parameters (e.g., density n,
temperature T ) but whose particle positions vary randomly from one realization to the next
see Section A.5.

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

from the value given in (2.2) for distances larger than the Debye length (cf.,
Fig. ??):
bmax = D .

(2.8)

To estimate the minimum impact parameter bmin , we note that when the
Coulomb potential energy qe qi /({4%0 }|x|) becomes as large as the electron
kinetic energy me v 2 /2: v$ becomes comparable to |v |, the scattering angle
becomes 90o [see Eq. (??)], and our weak interaction approximations
( break
(
down. Hence, we determine a classical minimum impact parameter by (v$ ( =
|v |, which yields
bcl
min =

Zi e2
Zi
Zi e2
Zi
&
=
& 4.8 1010
m. (2.9)
2
2
{4%0 }(me v )
{4%0 }(3Te )
12ne De
Te

Here, we have approximated me v 2 /2 by 3 Te /2, which is appropriate for a thermal electron in a Maxwellian distribution [cf., (??)]. Quantum mechanical effects become important when they could induce scattering through an angle
of 90o , which occurs [for wave scattering processes see (??)]) when the distance of closest approach b is less than half the radian de Broglie wavelength
h/mv = h/(2mv). This physical process
yields a quantum-mechanical
h /2
&
minimum impact parameter3 (for v & vT e 2Te /me )
bqm
min

h
1
&
& 1.1 1010 1/2
m.
2me v
4me vT e
Te (eV)

(2.10)

The relevant minimum impact parameter bmin is the maximum of the classical
and quantum-mechanical minimum impact parameters. Quantum-mechanical
> 20 Z 2 eV.
effects dominate for Te
i
With these specifications of the limits of integration, the impact parameter
integral in (2.7) can be written as
ln

bmax

bmin

db
= ln
b

D
bmin

)
qm *
bmin = max bcl
min , bmin ,

Coulomb logarithm.

(2.11)

It is called the Coulomb logarithm because it represents the sum or cumulative


effects of all Coulomb collisions within a Debye sphere for impact parameters
ranging from bmin to D .
To determine the relative magnitude and scaling of Coulomb collision effects,
it is convenient to assume classical effects determine the minimum impact parameter. When classical effects dominate (bmin = bcl
min ), the Coulomb logarithm
3 In Chapter 13 of Jacksons Classical Electrodynamics the factor of 2 is omitted in the
definition of the quantum-mechanical minimum impact parameter, but then the argument
of the Coulomb logarithm in (2.11) is multiplied by a factor of 2 when quantum-mechanical
effects dominate.

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS


becomes
ln = ln
cl

D
bcl
min

& ln

12 ne 3De
Zi

(2.12)

Since the definition of a plasma (cf., Section 1.8) requires that ne 3De >>>> 1,
plasmas have ln cl >> 1. For example, typical magnetic fusion experiments in
laboratory plasmas have n3D 106 , and hence ln 17.
Having defined the impact parameter integral in (2.7), the total Coulomb
collisional force on a test electron along its direction of motion thus becomes
,
+
dv$
)v$ *
4ni Zi2 e4
= )F$ * = me
=
ln

me v$ = me v$ . (2.13)
me
dt
t
{4%0 }2 m2e v 3
The Coulomb collisional drag force in the last form of this equation is called the
dynamical friction force because it is proportional to the test particle velocity.
Here, we have defined a net momentum loss or slowing down4 Coulomb collision
frequency for a particle of speed v in the Lorentz collision model:
(v)

4 ne Zi e4 ln
ln (12 ne 3De /Zi )
& pe
2
2
3
{4%0 } me v
4 ne 3De /Zi

Te
me v 2

$3/2
,

Lorentz collision frequency.

(2.14)

Here, compared to the form given in the next to last form in (2.13), we have
taken into account the condition for quasineutrality in a plasma: ne = Zi ni .
Note from the last form in (2.14) that the electron collision frequency is smaller
than the electron plasma frequency by a very large factor [ 1/(ne 3De ), which
is by definition a small number in a plasma]. The Lorentz collision frequency
2
can also be shown to be given by (v) = ni m v in which m = 4 (bcl
min ) ln
is a momentum transfer cross-section see Problems 2.6, 2.7. It can also be
deduced from the Langevin equation in which the stochastic force is due to
Coulomb collisions see Problem 2.8.
For classical hard collisions with b < bcl
min , the maximum parallel momentum transfer is given by max(v$ ) = 2v. The collision frequency for hard
2
collisions can be estimated using a cross section of hard & (bcl
min ) : hard =
2
)
,
which
is
smaller
than
the
collision
frequency
ni hard max(v$ ) & 2ni v(bcl
min
in (2.14) by a factor of 1/(2 ln ) << 1. Thus, the net Coulomb collision frictional force is dominated by the cumulative small angle collisions with impact
parameters b ranging between bmin and D that are embodied in the ln integral in (2.11). That is, the Coulomb logarithm represents the degree to which
cumulative small-angle collisions dominate over hard collisions for the Coulomb
collision processes in plasmas.
Detailed treatments of the physical phenomena of hard collisions for b bmin
(see Problems 2.7, 2.24) and of the Debye shielding process (see Chapter 13) for
4 Note that in the Lorentz collision model there is no energy transfer and only loss of
directed momentum see Problem 2.4 It is thus unfortunate and rather misleading that the
Lorentz collision frequency is often called a slowing down frequency in plasma physics.

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

> bmax = D yield order unity corrections to the ln ln (bmax /bmin ) factor
b
in (2.14). However, because these corrections are small and quite complicated,
it is customary to neglect them in most plasma physics calculations. Thus,
the Coulomb collision momentum loss frequency given in (2.14) and the other
Coulomb collision processes calculated in this chapter should be assumed to be
accurate to within factors of order 1/(ln ) 5 10%; evaluation of Coulomb
collision processes and their effects to greater accuracy is unwarranted.
Finally, we use our result for the Coulomb collision frictional force )F$ * on a
single electron to calculate the net frictional force on a drifting Maxwellian distribution of electrons flowing slowly (compared to their thermal speed) through
ez
a background of fixed, immobile ions. For a small net flow speed V$ in the
direction, the appropriate flow-shifted Maxwellian distribution for electrons is5
fM e (v)

#
$
me |v V$
ez |2
exp
2Te
,
#
$3/2
+
me v$ V$
2
me
eme v /2Te 1 +
+
& ne
2Te
Te
+
,
2
v 2 /vT
e
2 v$ V$
ne e
1 + 2 + ,
=
vT e
3/2 vT3 e
= ne

me
2Te

$3/2

(2.15)

in which in the last form


& we have used the convenient definition of the electron
thermal speed vT e 2Te /me . Multiplying (2.13) by this distribution and
integrating over the relevant spherical velocity space (v$ v = v cos ), the
Maxwellian-average (indicated by the bar over F$ ) of the Coulomb collisional
frictional force density on the drifting electron fluid becomes
%
d3v fM e (v) )F$ *
ne )F$ *

= e me ne V$ .

v 2 dv (v) me v

2vV$ ne ev /vT e
vT2 e
3/2 vT3 e

(2.16)

Here, we have defined the Maxwellian-averaged electron-ion collision frequency

#
$
4
4 2 ni Zi2 e4 ln
5 1011 ne Zi ln 1
s ,
e (vT e ) =
&
1/2 3/2
17
3
[Te (eV)]3/2
{4%0 }2 3 me Te

fundamental electron collision frequency. (2.17)

This is the average momentum relaxation rate for the slowly flowing Maxwellian
distribution of electrons. Since many transport processes arise from collisional
relaxations of flows in a plasma, this average or reference electron collision
5 Here, and throughout this text, a capital letter V (V) will indicate the average flow speed
(velocity) of an entire species of particles while a small letter v (v) will indicate the speed
(velocity) of a particular particle, or a particular position in velocity space.

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

10

CHAPTER 2. COULOMB COLLISIONS

frequency is often the fundamental collision frequency that arises in the


plasma electrical conductivity (see Section 2.3 below) and plasma transport
studies (cf., Problem 2.10).
Since a typical, thermal electron moves at the thermal speed vT e , it is convenient to define the characteristic length scale over which the momentum in a
flowing distribution of electrons is damped away by
#
$
2
vT e
17
16 [Te ((eV )]
& 1.2 10
, electron collision length. (2.18)
e
e
ne Zi
ln
Note that (in contrast to neutral particle collisions) it is not appropriate to
call this length a collision mean free path because a very large number
of random small-angle Coulomb collisions deflect particles velocities and cause
the net momentum loss over this length scale. [The total number of collisions
involved is of order n3D as a test electron traverses a Debye length times a
factor of e /D & n3D / ln , or of order (n3D )2 / ln >>>>>> 1.] For the
relevant length and time scales in some typical plasmas, see Problems 2.12.3

2.2

Diffusive Properties of Coulomb Collisions

The Coulomb collision process causes more than just momentum loss by the electrons. As indicated in Fig. 2.4, the dominant collisional process in individual
collisions is deflection of the test particle velocity in a random direction perpendicular to the original direction of motion. The net perpendicular Coulomb
collision force defined analogously to the net parallel force in (2.7) vanishes:
)v *
= ni v
)F * m
t

bmax

bmin

b db me v = 0.

While the ensemble average perpendicular force vanishes, because of the randomness of the impact angle , velocity-space deflections caused by Coulomb
collisions do have an effect in the perpendicular direction. Namely, they lead
to diffusion of the test particle velocity v. For a general discussion of diffusive
processes see Section A.5.
The temporal evolution of the velocity of a test particle as it undergoes random Coulomb collisions with background ions is illustrated in Fig. 2.5. While for
long times (many Coulomb collisions) the average of the perpendicular velocity
component vx vanishes ()vx * = 0), its square and the reduction of the velocity
component in the original direction of motion increase approximately linearly
with time )vx2 * & ()vx2 */t) t [see Eq. (??)] and v0 vz ()v$ */t) t.
The fact that the average of vx2 increases linearly with time while the average
of vx vanishes indicates a diffusive process for the x (perpendicular) component
of the test particle velocity. Because there is no preferred direction in the plane
2
*/2
perpendicular to the original direction of motion, we have )vx2 * = )vy2 * = )v
2
= (1/2)()v
*/t) t; hence there is velocity diffusion equally in both the x and
y directions.
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

11

Figure 2.5: Temporal evolution of vx , vx2 and vz components of the test particle
velocity as it undergoes random Coulomb collisions with background ions. Note
that for times long compared to an individual Coulomb collision time the average
of vx vanishes, but vx2 and v0 vz increase approximately linearly with time t.
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

12

CHAPTER 2. COULOMB COLLISIONS

To mathematically describe the diffusion in velocity space, we calculate the


mean square deflection of the test electron as it moves through the background
ions by the same ensemble-averaging procedure as that used in obtaining the
average parallel force in (2.7). We obtain
2
*
)v
ni v
t

bmax

bmin

b db v v =

8 ni Zi2 e4
ln = 2 v 2 .
{4%0 }2 m2 v
(2.19)

Thus, as can be inferred from (2.5), and from Figs. 2.4 and 2.5, in the Lorentz
2
*/v 2 t) for the test electron
scattering model the rate of velocity diffusion ()v
is twice the rate of momentum loss ()v$ */vt). Note that for the collisional
process being considered the velocity diffusion takes place at constant energy
and in directions perpendicular to the test particle velocity v; there is no speed
(energy) diffusion in the Lorentz collision model because the background particles (ions here) are assumed to be immobile and hence to not exchange energy
with the test electron.
In the spherical velocity space we are using, the pitch-angle through
which the random-scattering, deflections and diffusion take place is defined
by sin v /v = vx2 + vy2 /v. Since the Coulomb collision process is a smallangle random walk or diffusion process, the time required to diffuse the test
particle velocity vector through a small angle & v /v << 1 is much less than
the Lorentz collision model (momentum loss) time 1/, which is effectively the
time scale for scattering through 90 see Problem 2.12 for a specific example.
From (2.19) we can infer that collisional scattering through an angle << 1
(but must be greater than the for any individual Coulomb interaction so
a diffusive description applies) occurs in a time [see Fig. 2.5 and (??)]
t (v /v)2 / 2 / << 1/,

time to diffuse through << 1.

(2.20)

As time progresses, a test particles pitch-angle in velocity space is randomly


deflected or scattered. Thus, over time the pitch-angle
of a test particle assumes

a probability distribution that is of width t.


It can be shown (see Section 11.2) that the probability distribution for a test
particle with an initial velocity v0 [f (v, t = 0) (v v0 )] is given for short
times by
/
/
.
.
2
2
2
(v v0 ) e v /(2v t)
(v v0 ) e /2t
=
for t << 1.
ft (v, , t) &
2v02
t
2v02
t
(2.21)

'

This distribution function is normalized to represent one test particle: d3vft =


1. The delta function in speed, (v v0 ), represents the fact that the test
particle speed stays constant at the initial speed |v0 | v0 because the test
particle energy (speed) is constant in the Lorentz collision model. The factor
2
e /2t /(t) represents the diffusion in pitch-angle that takes place in a time
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

13

CHAPTER 2. COULOMB COLLISIONS

t; it indicates that ft is reduced by a factor of e1/2 & 0.61 for diffusion over
a pitch-angle of << 1 in the short time t 2 / indicated in (2.20). The
velocity-space diffusion properties of Coulomb collision processes are explored
in greater detail in Chapter 11.
The dynamical friction and diffusion coefficients for the Lorentz collision
model can be written in a coordinate-independent, vectorial form as follows.
First, note that the parallel or z direction here is defined to be in the initial
electron velocity direction:
ez v/v. Thus, we can write the dynamical friction
force coefficient due to Coulomb collisions in the form
)v$ *
)v*
=

ez = (v) v.
t
t

(2.22)

Similarly, because velocity diffusion occurs equally in all directions perpendicu2


*/t; hence the (second
lar to v, we have )vx2 */t = )vy2 */t = (1/2))v
rank tensor) diffusion coefficient can be written as
)v v*
t

2
)vy2 *
)vx2 *
1 )v
*

ex

ey
(
ex
ex +
ey =
ex +
ey
ey )
t
t
2 t
(2.23)
= (v) (v 2 I vv)

in which I is the identity tensor [see Eq. (??)]. These forms for )v*/t and
)vv*/t will be useful in Section 11.1 where we will develop a Lorentz
Coulomb collision operator for use in plasma kinetic theory.

2.3

Runaway Electrons and Plasma Resistivity

Next, we consider the combined effects of a macroscopic electric field E and the
dynamical friction due to Coulomb collisions on test electrons in a plasma. Using
the dynamical friction force given in (2.13) using the vectorial form indicated
in (2.22), Newtons second law for this situation can be written in the form
me

dv
= qe E me v.
dt

(2.24)

The electric field may be externally imposed, or arise from a collective response
in the plasma. The electric field E, which we take to be in the
ez direcez direction; Coulomb
tion accelerates electrons (qe = e) in the E or +
collisions exert a dynamical friction force that opposes this acceleration. In
a more complete Coulomb collision model that includes electron-electron collisions (see Section 2.7 below), the Lorentz collision frequency gets replaced
by a slowing down (subscript S) electron (momentum relaxation) collision
e/e
e/i
e/e
e/i
frequency Se = S + S , in which S and S are the momentum loss rates
for electron-electron and electron-ion collisions, which will be derived explictly
below, in Section 2.8. For electron-ion collisions, since electron speeds are typically much greater than the ion thermal speed and little energy is transferred
during the collisions because of the large disparity in masses, the ions are essentially immobile during the Coulomb collision process. Thus, the Lorentz
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

14

CHAPTER 2. COULOMB COLLISIONS

collision model is applicable and the relevant electron-ion collision frequency is


e/i
simply the Lorentz collision frequency: S = (v), as given in (2.14). Electronelectron collisions are in general more complicated because during collisions
both particles are in motion and energy is transferred. With these simplifications and adaptations, the equation governing the velocity of a single electron
in the
ez direction, (2.24) can be rewritten in the more precise one-dimensional
form
me

dv$
e/e
e/i
= (e)(E) (S + S ) me v$ = eE Se me v$ .
dt

(2.25)

We first consider the combined electric field and Coulomb collision effects
on energetic test electrons in the high energy tail of a Maxwellian distribution.
For these energetic test electrons the background electrons can be considered
at rest and the electron-electron momentum loss collision frequency is simply
e/e
S = 2 (v)/Zi the factor of two comes from the inverse dependence on
the reduced electron rest mass [see (2.56) below] and the 1/Zi factor eliminates
the dependence on the ion Zi in the Lorentz model collision frequency. The
total momentum loss collision frequency for these energetic electrons can thus
be written as
Se = (1 + 2/Zi ) (v) = (1 + 2/Zi ) (vT e ) vT3 e /v 3

for v >> vT e .

(2.26)

Here, the unity multiplicative factor (on ) represents electron-ion collisions


and the 2/Zi factor represents electron-electron collisions. Note that in the
limit Zi >> 2 this overall electron momentum relaxation rate becomes simply
the Lorentz model collision frequency and electron-electron collision effects are
negligible.
The dynamical friction force Se (v)me v$ in (2.25) with the Se given in (2.26)
decreases as v 2 for electrons in the high energy tail of a Maxwellian distribution. The dependence of the electric field and dynamical friction forces on the
speed v of a tail electron are illustrated in Fig. 2.6. As indicated, when the
electric field force exceeds the dynamical friction force, electrons are freely accelerated by the electric field. Such electrons are called runaway electrons. The
energy range for which runaway electrons occur is determined by eE > Se me v:
ED
me v 2
,
> (2 + Zi )
2 Te
|E|

where
ED

e ( 12 ln )
2ne e3 ln
me vT e (vT e )
=
=
,
2
{4%0 } Te
{4%0 }2De
eZi

(2.27)

Dreicer field

(2.28)

is a critical electric field strength, which, as indicated, is often called the Dreicer
field.6 For weak electric fields (|E| << ED ), the energy at which electron
6 H. Dreicer, Proceedings of the Second United Nations International Conference on the
Peaceful Use of Atomic Energy (United Nations, Geneva, 1958), Vol. 31, p. 57. See also,
Phys. Rev. 115, 238 (1959).

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

15

CHAPTER 2. COULOMB COLLISIONS

Force
runaway
electrons
e|E|

v
Figure 2.6: Relative strengths of the electric field e|E| and dynamical friction
Se (v)me v forces on an electron as a function of the electron speed v. Runaway
electrons occur when the electric field force exceeds the dynamical friction force.

runaways occur is far out on the high energy tail of the Maxwellian electron
distribution and only an exponentially small fraction of electrons run away see
Problem 2.13. [For relativistic electron energies the dynamical friction decreases
less rapidly than 1/v 2 and no runaways are produced for a weak electric field
satisfying |E|/ED < 2Te /(me c2 ) see Problem 2.14.] High Zi ions increase the
energy for electron runaway relative to that for protons because they increase
the frictional drag due to Coulomb collisions. Note also from the middle form
of the critical electric field defined in (2.28) that its magnitude is roughly (to
within a factor of 12 ln 10) what is required to substantially distort the
Coulomb electric field around a given ion [cf., (2.2)] at distances of order the
Debye length. Alternatively, it can be seen from the last form in (2.28) that
the Dreicer field is approximately the electric field strength at which typical,
thermal energy electrons with v vT e in a Maxwellian distribution become
runaways see Problem 2.15 for a more precise estimate. Thus, when the
electric field is larger than the Dreicer field, the entire distribution of electrons
responds primarily to the electric field and collisional effects are small.
For weak electric fields |E| << ED , most of the plasma electrons will be only
slightly accelerated by the E field before Coulomb collisions relax the momentum
they gain. However, the velocity distribution of electrons will acquire a net flow
Ve in response to the E field. Since the electric field has much less effect on
the more massive ions, the plasma ions will acquire a much smaller [by a factor
< 1/43 << 1] flow, which can be neglected. Thus, the electron
(me /mi )1/2
flow in response to the electric field will correspond to an electric current flowing
in the plasma. The proportionality constant between the current and the electric
field is the plasma electrical conductivity, which we will now determine.
For electrons with a flow-shifted Maxwellian distribution as in (2.15) that
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

16

CHAPTER 2. COULOMB COLLISIONS

ez Ve Vi ), the average (over


have a flow velocity Ve relative to the ions (V$
the Maxwellian distribution) frictional force is given in (2.16). Adding electric
field force and electron inertia effects yields the electron momentum density
equation
m e ne

dVe
= ene E me ne e (Ve Vi ),
dt

(2.29)

in which e is the fundamental electron collision frequency defined in (2.17). In


equilibrium (t >> 1/e , d/dt 0) we obtain the current induced by an electric
field:
J = ne e(Ve Vi ) = 0 E,

Ohms law

(2.30)

in which
0 =

ne e2
1
,
me e

reference (subscript 0) plasma electrical conductivity,


(2.31)

where is the plasma resistivity. The electron collision frequency that enters
this formula is e , which is the (electron-ion) Lorentz collision frequency (2.14)
averaged over a flowing Maxwellian distribution of electrons given in (2.17). (In
this analysis the electron Coulomb collision frequency is assumed to be much
greater than the electron-neutral collision frequency. See Problems 2.19, 2.20
for situations where this assumption is not valid and the electrical conductivity
3/2
is modified.) Note also that since ne /e Te , the electrical conductivity in a
3/2
plasma increases as Te an inverse dependence compared to solid conductors
whose electrical conductivity decreases with temperature. The conductivity in
plasmas increases with electron temperature because the noise level [see (??)]
and collision frequency [see (2.17)] decrease with increasing electron temperature
and Debye length. For some perspectives on the magnitude and effects of the
electrical conductivity in plasmas, see Problems 2.162.18.
In a more complete, kinetic analysis with the Lorentz collision model (see
Section 11.4), the electric field distorts the electron distribution function more
than indicated by the simple flow effect in (2.15). Specifically, we can infer
from (2.25) and (2.26) that higher energy electrons receive larger momentum
input from the electric field because the Coulomb collision dynamical friction
force decreases as v 2 . Thus, the current is carried mainly by higher energy
(v 2 vT e ), more collisionless electrons than is embodied in the simple flowshifted Maxwellian distribution. Since the collision frequency decreases as 1/v 3 ,
the average collision frequency is reduced, by a factor of 3/32 & 0.2945 e ;
hence the electrical conductivity in a kinetic Lorentz collision model is increased
relative to that given in (2.31) by the factor 1/e .
Electron-electron collisions are momentum conserving for the electron distribution function as a whole. Thus, they do not contribute directly to the
momentum loss process or the plasma electrical conductivity. However, since in
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

17

CHAPTER 2. COULOMB COLLISIONS

a kinetic description the electric field distorts the electron distribution function
away from a flow-shifted Maxwellian, electron-electron collisions have an indirect effect of reducing the net flow (and electrical conductivity) in response to
the electric field as they try to force the electron distribution to be as close
to a Maxwellian as possible. Details of this process will be discussed in Section
12.3.
The net result of these kinetic and electron-electron effects, which is obtained
from a complete, kinetic analysis that was first solved numerically by Spitzer
and H
arm,7 is that the effective electron collision frequency is reduced by a
generalized factor e . Thus, the electrical conductivity becomes
Sp =

ne e2
0
=
,
me e e
e

Spitzer electrical conductivity.

(2.32)

The generalized factor e ranges from 0.5129 for Zi = 1 to 3/32 & 0.2945
for Zi (Lorentz kinetic model). A later analytic fluid moment analysis8
has shown that this factor can be approximated to three significant figures (see
Section 12.3?), which is much more accuracy than warranted by the intrinsic
< 10%) of the Coulomb collision operator, by
accuracy ( 1/ ln
e &

2.4

1 + 1.198Zi + 0.222Zi2
.
1 + 2.966Zi + 0.753Zi2

(2.33)

Effects of Coulomb Collisions

So far we have concentrated on the electron momentum relaxation effects of


Coulomb collisions using a Lorentz collision model. In this section we discuss
phenomenologically more general Coulomb collision effects on electrons as well
as the collisonal effects on ions, and between ions and electrons. A complete,
rigorous treatment of Coulomb collision effects begins in Section 2.6.
The Lorentz collision model takes into account electron-ion collisions but
neglects electron-electron collisions. However, these two collisional processes
occur on approximately the same time scale, at least for ions with a Zi that is
not too large. As indicated in the preceding section, electron-electron collisions
tend to relax the electron velocity distribution toward a Maxwellian distribution function. They do so on approximately the fundamental electron collision
time scale 1/e . However, as indicated in (2.25) and (2.26), the relaxation of
electrons in the high energy tail of the distribution is slower. The characteristic time for tail electrons to equilibrate toward a Maxwellian distribution is
(v/vT e )3 /e for v >> vT e . (For an application where this effect is impor< vT e relax toward a
tant see Problem 2.21.) In contrast, all electrons with v
Maxwellian distribution on the same time scale as the bulk: 1/e .
7 L.

Spitzer and R. H
arm, Phys. Rev. 89, 977 (1953).
Hirshman, Phys. Fluids 20, 589 (1977).

8 S.P.

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

18

CHAPTER 2. COULOMB COLLISIONS

As indicated in (2.16) and (2.29), the net Coulomb collisional force density
on a Maxwellian distribution of electrons flowing relative to the ions is
ne e J
,

Re me ne e (Ve Vi ) =

collisional friction force density.


(2.34)

This is the electron force density that was introduced in the electron fluid momentum balance given in (2.29). Note also from a temporal solution of (2.29)
that the electron flow (momentum) will relax exponentially to its equilibrium
value at the rate e , i.e., on the electron time scale e = 1/e . Because Coulomb
collisions are momentum conserving, any momentum lost from the electrons
must be gained by the ions. Thus, the Coulomb collisional force density on ions
is given by
Ri = Re .

(2.35)

Ion-ion collisions are analogous to electron-electron collisions and complicated during Coulomb collisions both particles are in motion and energy is
exchanged between them. Nonetheless, considering a Lorentz-type model for
ion-ion collisions using the framework developed in Section 2.1, it is easy to
see that the appropriate ion collision frequency should scale inversely with the
square of the ion mass and the cube of the ion speed. A detailed analysis (see
Sections 2.62.8) of the effects of ion-ion collisions yields a flowing-Maxwellianaveraged ion collision frequency given by
i =

4 ni Zi4 e4 ln
1/2

3/2

{4%0 }2 3 mi Ti

me
mi

$1/2 #

Te
Ti

$3/2

Z2
i e ,
2

fundamental ion collision frequency.

(2.36)

The 2 factor (in the denominator at the end of the second formula) enters because of the combined effects of the reduced mass [see (2.56) below]
and the motion of both particles during ion-ion collisions. Note that for an
electron-proton (Zi = 1) plasma with Te Ti the ion collision frequency is
smaller than the electron collision frequency by a square root of the mass ratio:
< 1/60 << 1. Because of their very disparate masses,
i /e (me /2mi )1/2
ion-electron collisional effects are typically smaller than ion-ion collisional ef< 1/43 << 1; hence they are negligible for ion
fects by a factor of (me /mi )1/2
collisional effects. As for electrons, ion collisions drive the velocity distribution of ions toward a Maxwellian distribution on the ion collisional time scale
i = 1/i (2mi /me )1/2 /e >> 1/e . In addition, like electrons, ions in the
high energy tail of the distribution relax toward a Maxwellian distribution more
slowly: (v/vT i )3 /i for v >> vT i .
We now determine the small energy transfer from electrons to ions during
Coulomb collisions which we have heretofore neglected. Momentum is conserved
during a Coulomb collision. Thus, if an electron acquires an impulse me ve
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

19

CHAPTER 2. COULOMB COLLISIONS

during a electron-ion collision, the ion acquires an impulse determined from


momentum conservation:
me ve + mi vi = 0

vi = (me /mi ) ve .

The energy exchange from electrons to ions initially at rest during a Coulomb
collision will thus be
#
$2
#
$
mi me
me me
mi
2
vi vi =
v
ve ve &
.
2
2
mi
mi
2
The net energy exchange from a test electron moving through the backgound
stationary ions can thus be evaluated using (2.19):
#
$
me
mi )vi vi *
=
me v 2 (v).
(2.37)
2
t
mi
Note that this energy exchange rate is slower than the basic Lorentz collision
< 103 << 1, due to the large disparity
frequency by a factor of me /mi
between the electron and ion masses.
Integrating this last result over a Maxwellian distribution of the electrons,
the Maxwellian-averaged rate of energy () density transfer from electrons to
e/i
initially stationary background ions ( in Section 2.7) becomes
%
me
mi )vi vi *
e/i
ne Te = 3
e ne Te .
d3v fM e
2
t
mi

A more complete analysis (see Section 2.8) shows that if the background ions
have a Maxwellian velocity distribution (instead of being stationary and immobile as they are in the Lorentz model) Te Te Ti in this formula, as would be
expected physically. Thus, the rate of ion energy density increase from Coulomb
collisions with electrons is
me
e/i
Qi ne (Te Ti ) = 3
e ne (Te Ti ),
mi

ion collisional heating density.


(2.38)

In the absence of other effects, the equation governing ion temperature evolution
[see (??)] becomes
me
3 dTi
ni
= Qi = 3
e ne (Te Ti ).
2
dt
mi

(2.39)

Here, (3/2)(ni dTi /dt) represents the rate of increase of ion internal energy in the
plasma. From (2.39) we see that for a constant electron temperature the characteristic time scale on which Coulomb collisions equilibrate the ion temperature
e/i
> 103 /e >>
to the electron temperature is ie = 3/(2 ) = (mi /2me )/e
1/e . (For a more precise determination of the temporal evolution of the collisional equilibration of the electron and ion temperatures in a plasma, see Problem 2.23.)
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

20

CHAPTER 2. COULOMB COLLISIONS

Because energy is conserved in the elastic Coulomb collisions, energy gained


by the ions is lost from the electrons. In addition, the electrons are heated by the
work they do per unit time in flowing relative to the ions against the collisional
friction force density Re given in (2.34). Thus, the total electron heating due
to Coulomb collisions is given by
Qe = (Ve Vi ) Re Qi = J 2 / Qi .

(2.40)

In the absence of other effects and using 1/ = , the electron temperature


evolution equation becomes
3 dTe
ne
= Qe = J 2 Qi .
2
dt

(2.41)

In these equations J 2 is the joule or ohmic heating induced by a current density


J flowing in a plasma with resistivity . Because the plasma resistivity scales
3/2
, for a constant current density the joule heating rate of a plasma
as Te
decreases as it is heated. Thus, joule heating becomes less effective as the
electron temperature increases. Note also that despite the complexity of the
dependence of e on Te , the characteristic time scale for Coulomb collisional
relaxation of the electron temperature to its equilibrium value is roughly the
same as the ion temperature equilibration time scale ie .

2.5

Numerical Example of Collisional Effects*

In order to illustrate the evaluation of and numerical values for these various
collisional processes, we will work them out for a particular plasma example.
The plasma example will be chosen to be typical of laboratory experiments
for magnetic fusion studies, but the plasma will be assumed to be infinite and
uniform, and in equilibrium so there will not be any spatial or temporal
inhomogeneity effects. For the plasma electrons we assume an electron density
ne = 2 1019 m3 and electron temperature Te = 1 keV. For these parameters
the electron plasma period [inverse of electron plasma frequency from (??)] is
1/pe = 1/[56(21019 )1/2 ] & 41012 s1 , the electron Debye length from (??)
is De = 7434[103 /(2 1019 )]1/2 & 5.3 105 m, and the number of electrons
in an electron Debye cube is ne 3De & 3 106 . These parameters clearly satisfy
the criterion ne 3De >> 1 for the plasma state.
The ions in laboratory plasmas often include impurities in addition to the
desired hydrogenic species. We will take into account an impurity species to
show how the various plasma collision rates presented in the preceding sections
need to be modified to take into account multiple species of ions, and in particular impurities. For our example laboratory plasma we will assume a dominant
deuterium (atomic weight AD = 2, charge ZD = 1) ion species with relative
density nD /ne = 0.64 and fully ionized carbon (AC = 12, ZC = 6) impurities
with a relative density of nC /ne = 0.06. Note that even though the carbon ion
density is only 6% of the electron density the carbon ions supply 36% of the ion
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

21

CHAPTER 2. COULOMB COLLISIONS

0
charge needed for charge neutrality: i ni Zi = [0.64 + (0.06)(6)]ne = ne . Both
the deuterium and carbon ion temperatures will be assumed to be 0.5 keV.
In order to calculate the ln factor for the fundamental electron collision rate
we first need to determine the maximum and minimum collisional impact parameters bmax and bmin . The maximum impact parameter is the overall plasma
Debye length in the plasma which is defined in (??). For our multi-species
plasma the plasma Debye length can be calculated from the electron Debye
length by taking out common factors in the ratio of D to De :
D De

2 ns T e
s

ne Ts

Zs2

31/2

(2.42)

which for our plasma yields D = De /[1 + (0.64)(2) + (0.06)(2)(62 )]1/2 &
De /2.6 & 2 105 m. Classical and quantum mechanical minimum impact parameters for electron-deuteron collisions in this plasma are estimated
10
/103 = 1.4 1012 m and bqm
from (2.9) and (2.10): bcl
min = 4.8 10
min =
10
3 1/2
12
= 3.5 10
m. Since the quantum mechanical impact
1.1 10 /(10 )
parameter is larger, we use it for bmin and thus have ln ln(D /bqm
min ) &
ln[(2 105 )/(3.5 1012 )] & ln(5.7 106 ) & 16. Since the Coulomb collision
frequency is only accurate to order 1/ ln & 1/16 & 0.06, in the following we
will give numerical values to only about 6% accuracy; more accuracy is unwarranted and misleading.
In calculating the electron collision frequency we need to take account of all
the ion species. From (2.17) we see that the electron-ion collision frequency is
proportional to ni Zi2 . Thus, for impure plasmas it is convenient to define
Zeff

0
0
2
2
i ni Zi
i ni Zi
0

=
,
ne
i ni Zi

effective ion charge

(2.43)

in which the sum is over all ion species in the plasma. Hereafter in this section,
we will designate the main ions with a subscript i and the impurities with
a subscript Z. Then, for our example plasma Zeff (ni Zi2 + nZ ZZ2 )/ne =
0.64(12 ) + (0.06)(62 ) = 2.8. The overall electron collision frequency e defined
in (2.17) for an electron-ion plasma can be written for an impure plasma in
terms of the electron-deuterium (dominant ion with Zi = 1) collision frequency
e/Z =1
e i as follows:
e = Zeff ee/Zi =1 ,
in which
ee/Zi =1

4 2 ne e4 ln
{4%0

}2

1/2 3/2
3 me Te

&

5 1011 ne
[Te (eV)]3/2

(2.44)
#

ln
17

s1 .

(2.45)

e/Z =1

& (5 1011 )(2 1019 )(16/17)/(103 )3/2 &


For our example plasma e i
3 104 s1 , which gives e = (2.8)(3 104 ) & 8.4 104 . Hence, for our
example plasma the time scale on which the electron distribution becomes a
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

22

CHAPTER 2. COULOMB COLLISIONS

Maxwellian and electron flows come into equilibrium is e 1/e & 12 s.


The distance typical electrons travel in this time is the electron collision length
(2.18) e vT e /e , which is about 230 m for our plasma. Finally, the reference
electrical resistivity calculated from (2.31) is about 1.5 107 m. For impure
plasmas it is appropriate to replace the Zi in (2.33) by Zeff , which then yields
e & 0.4 for Zeff & 2.8. Thus, the Spitzer electrical resistivity for our example
plasma becomes 6 108 m. For reference, the resistivity of copper at room
temperature is about 1.7 108 m, a factor of about 3.5 smaller.
To calculate the ion collision frequency for the dominant ions (subscript i)
in an impure plasma we need to include both their self-collisions and their collisions with impurities (subscript Z). Since the masses of impurity ions are rather
disparate from the
dominant ions (mi << mZ AD << AC for our example plasma), the 2 rest mass factor is not appropriate for collisions between
dominant ions and impurities. Thus, the appropriate collision frequency for the
dominant ions in an impure plasama becomes
i/i

i = fi i
with
i/i
i

4 ni Zi4 e4 ln
1/2

3/2

{4%0 }2 3 mi Ti

fi 1 +

nZ ZZ2
ni Zi2

$#

mi
mZ

ni Zi4
ne

$1/2

$#

(2.46)

me
mi

$1/2 #

Te
Ti

$3/2

e/Zi =1

(2.47)

ion collisions impurity factor.

(2.48)

For multiple impurity species (Z) one just


sums the second term in fi over
+ 2 [(0.06)(62 )/0.64](2/12)1/2 & 3, and
them. For our example plasma fi = 1
i & 3(0.64)(1/3672)1/2 23/2 (3 104 )/ 2 & 1.9 103 s1 . Thus, the ions will
relax toward a Maxwellian distribution and their equilibrium flow on the ion
collisional time scale i = 1/i & 530 s. The ion collision length defined by
i = vT i /i is about 120 m for our plasma, which is about a factor of two less
than the electron collision length e .
Finally, we calculate the longest time scale process ion-electron energy
exchange. We must again take account of impurities in the calculation. Here,
since an electron-ion mass ratio is involved, we obtain
#
$#
$
me
ni Zi2
e/i
(2.49)
ee/Zi =1
= fie 3
mi
ne
in which the relevant factor to include impurity effects is
#
$#
$
nZ ZZ2
mi
, ion-electron energy exchange impurity factor.
fie = 1 +
ni
mZ
(2.50)
Again, for multiple impurity species (Z) one just sums the second term in fie
over them. For our example plasma fie = 1 + [(0.06)(62 )/(0.64)](2/12) &
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

23

CHAPTER 2. COULOMB COLLISIONS

1.6. In the presence of impurities the time scale for ion-electron temperature
e/i
equilibration becomes [see discussion after (2.39)] ie 3/(2 ), which for
4
our plasma is ie = (3672/2)/[(1.6)(0.64)(3 10 )] & 60 ms.
In summary, the electron, ion and ion-electron collision times in our example
plasma are e : i : ie & 12 : 530 : 60 000 s. Their ratios are in rough
accord with their anticipated mass ratio scalings of 1 : (mi /me )1/2 : mi /me
= 1 : 61 : 3672. Note also that all these collision time scales are much much
longer (by a factor ne 3De & 3 106 >> 1) than the plasma oscillation period
1/pe & 4 106 s.
Implicit in the preceding analysis is the assumption that no other physical
processes operate on the charged particles in the plasma on these character> 100 m)
istic collision time (& e ie 10 104 s) or length (& e , i
scales. In practice, in most plasmas many other processes (for example, temporal variations, gyromotion in magnetic fields, and spatial inhomogeneities)
vary more rapidly than one or more of these collisional effects and modify or
impede the collisional processes. Such combined collision and geometric effects
will be discussed later, particularly in Part IV: Transport. Note, however, that
even in the limit of very short time scales (compared to e ) Coulomb collision
effects are not insignificant; as indicated by (2.20), in a time t they diffusively
spread the velocity vectors of charged particles in a plasma through a pitch-angle
& v /v & (t)1/2 . This velocity diffusion effect is important in smoothing out
sharp gradients in velocity space and leads to collisional boundary layers in
otherwise collisionless plasmas. Thus, Coulomb collisions will often play a
significant role even in collisionless plasmas. In fact, as we will see in later
chapters, Coulomb collisions provide the fundamental irreversibility (entropy
producing relaxation mechanisms) in plasmas.

2.6

Collisions with a Moving Background+

The most general Coulomb collision processes are those where a test particle
species (s) collides with an arbitrary background species (s& ) of plasma particles
that are in motion, which we now consider. The test particle charge, mass,
position and velocity vectors will be taken to be qs , ms , x and v while the corresponding quantities for the background particles will be qs" , ms" , x& and v& . The
background particles will be assumed to have an arbitrary velocity distribution
given by fs" (v& ).
The procedure we follow to determine the Coulomb collision processes for
this general case follows that used in the Lorentz collision model except that
now the basic interaction is most conveniently calculated in a center-of-mass (or
really -momentum) frame. To develop the equations of motion in a center-ofmomentum frame, we first note that the equations of motion of the interacting
test and background particles are given by
ms

DRAFT 14:53
August 25, 2003

dv
qs qs" x x&
= qs E(x) =
,
dt
{4%0 } |x x& |3
c
!J.D
Callen,

(2.51)

Fundamentals of Plasma Physics

24

CHAPTER 2. COULOMB COLLISIONS


ms"

dv&
qs qs" x& x
= qs" E(x& ) =
.
dt
{4%0 } |x& x|3

(2.52)

Note that the forces in these equations are equal and opposite because of the
conservative nature of the Coulomb force. Defining the center-of-momentum
position and velocity vectors as
R=

ms x + ms" x&
,
ms + ms"

U=

ms v + ms" v&
,
ms + ms"

(2.53)

and the corresponding relative position and velocity vectors


u = v v& ,

r = x x& ,

(2.54)

we find the equations of motion in (2.51), (2.52) become


dU
= 0,
dt

mss"

q s q s" r
du
=
,
dt
{4%0 }|r|3

(2.55)

in which mss" is defined by


mss"

ms ms"
,
ms + ms"

reduced mass.

(2.56)

From the first relation in (2.55) we find that the center-of-momentum velocity U
is constant through the Coulomb collision interaction of the test and background
particles.
The equation describing the force on the relative velocity u v v& in
(2.55) is analogous to that in (2.2) for the Lorentz collision model. Adopting
a coordinate system analogous to that in Fig. 2.3 in which v is replaced by
u |v v& |, we readily find that the change u in a single Coulomb collision
interaction between a test particle (s) and background particle (s& ) is
%
1
q s q s" r
2qs qs"
dt
=
(
ex cos +
ey sin ) . (2.57)
u =
mss" {4%0 }r3
{4%0 }mss" bu
Since the total energy is constant in the center-of-momentum frame for an elastic Coulomb collision, using a geometry analogous to that in Fig. 2.4, with v
replaced by the relative velocity u, and relations (2.5), (2.6), we obtain
1
1
u u = u u & u u
2
2

u$ =

2qs2 qs2"
.
{4%0 }2 m2ss" b2 u3
(2.58)

Next, we want to determine the dynamical friction and diffusion coefficients


"
"
)v*s/s and )vv*s/s for test particles s colliding with background particles
s& . To do so we must relate v to the relative u determined above. Utilizing
the momentum conservation relations arising from U = constant in (2.53) with

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

25

CHAPTER 2. COULOMB COLLISIONS

v v + v, v& v& + v& and u u + u from before to after the collision,


we find
ms
mss"
v, v =
u.
(2.59)
v& =
m s"
ms
Then, taking account of the velocity distribution fs" (v& ) of the background particles, we define the average vectorial dynamical friction and tensorial velocity
diffusion coefficients to be
%
%
%
"
mss"
)v*s/s

d3v & fs" (v& ) u d b db


u,
(2.60)
t
ms
%
%
%
"
)vv*s/s
m2 "
3 &
&
"

d v fs (v ) u d b db ss2 uu.
(2.61)
t
ms
Using (2.57) and (2.58), the integrations in (2.60) and (2.61) can be performed with a specification of the impact parameter integral in (2.11) generalized to a test particle (s) colliding with a moving background (s& ) as follows:
#
$
% bmax
)
D
db
qm *
"
= ln
ln ss
, bmin = max bcl
(2.62)
min , bmin
b
b
min
bmin

in which

bcl
min

q s q s"
{4%0 }mss" u2

bqm
min =

h
4mss"

& .
u2

(2.63)

The u2 indicates an average of u2 over the distribution of background particles;


an appropriate typical value for this quantity is given in (2.99) below. In what
follows we will implicitly assume that ln ss" is independent of v& so that it can be
brought outside the v& integration in equations (2.60) and (2.61); retaining the
ln inside the v& integration would only yield negligible (additional) corrections
of order 1/ ln to the results we obtain below.
Thus, performing the integrations in (2.60) and (2.61) utilizing the impact
parameter integral in (2.62) and the facts that
u
u
= ,
v
u

1
u
= 3,
v u
u

u2 I uu
2u
1
=
ex
= (
ex +
ey
ey ) ,
3
vv
u
u

(2.64)

for our present velocity space coordinate system we obtain (for an alternate
derivation using the Rutherford differential scattering cross section see Problem 2.24):
#
$%
"
ms 4qs2 qs2" ln ss"
u
Hs" (v)
)v*s/s
=
, (2.65)
d3v & fs" (v& ) 3 ss"
t
mss"
{4%0 }2 m2s
u
v
#
$%
"
2
4qs2 qs2" ln ss"
2 Gs" (v)
)vv*s/s
3 &
& u I uu
" (v )
"
=
, (2.66)
v
f

d
s
ss
t
{4%0 }2 m2s
u3
vv
in which
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

26

CHAPTER 2. COULOMB COLLISIONS


ss"
Gs (v)

4qs2 qs2" ln ss"


,
{4%0 }2 m2s

d3v & fs" (v& ) |v v& |,


#
$%
%
&
&
ms
ms
3 & fs" (v )
3 & fs" (v )
=
1
+
v
.
dv
d
Hs" (v)
mss"
|v v& |
ms"
|v v& |

(2.67)
(2.68)
(2.69)

The G and H functions are formally similar to the electrostatic potential due to
a distributed charge 'density for which Poissons equation 2 = q (x)/%0 has
the solution (x) = d3x& q (x& )/({4%0 }|x x& |). They are called Rosenbluth
potentials9 . Using the facts that
#
$

1
21

= 4(u) = 4(v v& ),


(2.70)
v
u
v v u
4
5

u
2
u

= ,
(2.71)
2v u =
v v
v
u
u
the Rosenbluth potentials can be shown to satisfy the relations
2v Hs" (v) = 4(1 + ms /ms" )fs" (v),
2v Gs" (v) = 2Hs" (v)/(1 + ms /ms" ),

(2.72)

2v 2v Gs" (v) = 8fs" (v).

Note that since the second of these equations shows that Hs" is proportional to
a Laplacian velocity space derivative of Gs" , the Rosenbluth potential Gs" is the
fundamental one from which all needed quantities can be derived.
From the analogy of the first of the forms in (2.72) to electrostatics and
the definition of )v*/t in (2.65) in terms of the Rosenbluth potential Hs" ,
we see that the dynamical friction )v*/t tries to relax the test particle velocity to the centroid of the velocity distribution of the background particles
fs" (v) see Problems 2.25 and 2.26. However, the velocity space diffusion
)vv*/t causes the velocity distribution of the test particles to maintain a
thermal spread comparable to that of the background particles. The dynamical
balance between these two collisional processes on an entire distribution of test
particles determines their collisional distribution function see Chapter 11.
Finally, noting that
# 2
$
# $
u I uu
2u

1
2 u

2 u

=
2
= 2
, (2.73)
=
v
v
u3
v vv
v
v u
u v
we find that for Coulomb collisions the dynamical friction and velocity diffusion
coefficients are related by the important relation
#
$
"
"
"
ms )vv*s/s
1 + ms /ms"
)vv*s/s
)v*s/s
=

.
t
2mss" v
t
2
v
t
(2.74)
9 M.N.

Rosenbluth, W. MacDonald and D. Judd, Phys. Rev. 107, 1 (1957).

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

27

CHAPTER 2. COULOMB COLLISIONS

The total collisional effects on a test particle due to Coulomb collisions with
all types of background particles are obtained by simply adding the contributions
from each species of background particles:
2 )v*s/s
)v*s
=
,
t
t
"
s

"

2 )vv*s/s
)vv*s
=
.
t
t
"

(2.75)

Note also that the combination of this summation of species effects and, more
importantly, of the fact that the Rosenbluth potentials are integrals over the
background distribution functions, means that the dynamical friction and velocity diffusion coefficients are not sensitively dependent on detailed features of
fs" (v). (Recall the analogous weak dependence of an electrostatic potential to
the distribution of charges inside a surface.) Thus, evaluation of the Rosenbluth potentials for Maxwellian background distributions will be useful both
in describing test particle collisional processes in Maxwellian plasmas and in
other plasmas of interest where the distribution functions are reasonably close
to Maxwellians.

2.7

Collisions with a Maxwellian Background+

Specific test particle collisional effects due to dynamical friction and velocity
diffusion can be worked out in the rest frame of the background particles for an
isotropic Maxwellian velocity distribution of the background particles:
fM s" (v) = ns"

ms "
2Ts"

$3/2

ems" v

/2Ts"

ns" ev /vT s"


.
3/2 vT3 s"

(2.76)

Here, we have defined a typical thermal speed [see (??)]


vT s" (2Ts" /ms" )1/2 .

(2.77)

Note that this speed is not the average speed [see (??)] for a Maxwellian distribution, which is (8Ts" /ms" )1/2 ; however, it is the most probable speed [see
(??)] and it is mathematically convenient.
For a Maxwellian velocity distribution the Rosenbluth potential Gs" (v) defined in (2.68) can be evaluated in a spherical coordinate system in the relative
velocity space u = v v& as follows:

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

28

CHAPTER 2. COULOMB COLLISIONS

'
d3v & fs" (v& ) |v v& | = d3 u fs" (u + v) u
%
% 1
2
2
2
n s"
2
2u du
d(cos ) u e(v +u +2uv cos )/vT s"
= 3/2 3
vT s" 0
1
%
7
2
ns" vT2 s" u2 du 6 (v+u)2 /v2 "
(vu)2 /vT
Ts e
s"
=
e
v 0 vT3 s"
1
3
%
% x

ns" vT s" 1
y 2
2
y 2
4 x dy ye
2
dy (y + x)e
=
x

x
0
1
3
%
% x
% x

ns" vT s" 2
y 2
2 y 2
y 2
2 x dy ye
+
dy y e
+x
dy e
=
x

x
0
0
(2.78)

Gs" (v)

'

in which x v/vT s" . The integrals in the last forms of (2.78) are related
to the error function or probability integral (cf., Problem 2.27), but are most
conveniently written in terms of the Maxwell integral defined by
% x
2
dt t et ,
(2.79)
(x)
0
which has the properties
d
2 x
=
xe ,

dx

&

2
+ =

&

2
dy ey erf ( x ).

(2.80)

Physically, the Maxwell integral is the normalized integral of a Maxwellian velocity distribution out to a sphere of radius v. Utilizing these definitions, we
find that the Rosenbluth potential Gs" (v) for a Maxwellian distribution of background particles can be written as
1
Gs" (v) = ns" vT s" [(x + 1) & (x) + (x + 1/2)(x)] ,
x
in which
"

x xs/s =

ms" v 2
v2
= 2
2Ts"
vT s"

(2.81)

(2.82)

is the square of the ratio of the test particle speed to the thermal speed of the
background particles of species s& .
Thus, for an isotropic Maxwellian velocity distribution of background particles the Rosenbluth potential Gs" (v)=Gs" (v); that is, it depends only on the test
particle speed v, not its velocity v. Then, as can be shown from (2.66), )vv*
2
*/2 and )v$2 *. Furis a diagonal tensor with elements )vx2 * = )vy2 * = )v
ther, it can be shown that )v* is in the
ez or v direction. [These properties are
valid for any distribution function for which the Rosenbluth potential Gs" (v)

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

29

CHAPTER 2. COULOMB COLLISIONS

depends only on the test particle speed v.] Substituting the Rosenbluth potential in (2.81) into (2.65) and (2.66), and utilizing (2.72) or (2.74), we find that
the relevant dynamical friction and velocity diffusion coefficients are given by
"
+
#
$,
+
,
)v$ *s/s
ms
ms 1
s/s"
2 Gs"
= ss"
v
=

(x)
0 v, (2.83)
2
t
2mss" v v v
v
mss"
+
#
$
,
2 s/s"
1
2 Gs"
)v *
s/s"
&
= ss"
= 2 (x) 1
+ (x) 0 v 2 ,
(2.84)
t
v v
2x
"
+
,
)v$2 *s/s
(x) s/s" 2
2 G s"
= ss"
=
0 v .
(2.85)
t
v 2
x

1 0
Note that in contrast to the Lorentz collision model, we now find )v$2 * =
because the background particles are of finite mass and in motion, and hence
can exchange energy with the test particle during a Coulomb collision. The net
rate of change of the test particle energy, which is given by (m/2))v 2 */t
(m/2))(v + v) (v + v) v 2 */t, can be determined from these coefficients
as well:
"

"
2 s/s"
)v$2 *s/s
)v$ *s/s
)v
*
)v 2 *s/s
= 2v
+
+
t
t
t
t
s/s"
= 2 [(ms /ms" )(x) & (x)] 0 v 2 .

"

(2.86)

The reference collision frequency for all these processes is defined by


s/s"

ns" ss"
4 ns" qs2 qs2"
=
ln ss"
v3
{4%0 }2 m2s v 3
#
$
ns" Zs2 Zs2"
ln ss"
11 1
& (6.6 10 s )
, (2.87)
17
(ms /me )1/2 (Es /eV)3/2

0 (v)

which is a straightforward generalization of the collision frequency (v) derived


for the Lorentz collision model in (2.14).
These dynamical friction and velocity diffusion coefficients can be utilized to
elucidate the rates at which the various Coulomb collision processes affect the
test particle velocity. Thus, we define the rates for momentum loss or slowing
down (S ), perpendicular diffusion ( ), parallel or speed diffusion ($ ) and
energy loss ( ) resulting from collisions of a test particle s on a Maxwellian

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

30

CHAPTER 2. COULOMB COLLISIONS


velocity distribution of background particles s& as follows:10
d
)v*s/s
s/s"
(ms v) = S ms v ms
dt
t
d
|v v|2 =
dt

v 2

d
|v v|2$ =
dt

s/s"
$ v 2

"
d
s = s/s s
dt

s/s"

"

2 s/s
)v
*
t

)v$2 *ss

+#
$ ,
ms
s/s"
1+
0 ms v,
ms"
+
,

s/s"
= 2 + &
0 v 2 ,
2x
+ ,
s/s" 2
=
v ,
x 0
+
,
ms
s/s"
= 2
& 0 s
m s"
(2.88)
=

"

"

ms )v 2 *s/s
2
t

"

in which v is the average test particle velocity (see Section 2.10 for a detailed
specification of v), s ms v 2 /2 is the test particle energy, and |v v|2 ,
|v v|2$ indicate the diffusional spread of the test particle velocity in directions
perpendicular, parallel to its direction of motion. From the definitions in (2.86)
through (2.88) we see that is not an independent quantity:
= 2 S $ .

(2.89)

The total collisional effects due to all background particles are obtained by
summing over s& as indicated in (2.75). The overall picture of how these Coulomb
collision effects slow down and diffuse the test particle velocity are indicated
schematically in Fig. 2.7.
Equations (2.82)(2.88) provide a very complete and useful description of
the evolution of the velocity of a test particle of species s suffering Coulomb
collisions with Maxwellian background particles s& see Problems 2.282.33
for some illustrative applications of them. In addition, they can be used to
develop a Monte Carlo scattering operator for numerical studies of the effects of
Coulomb collisions on the velocity of a test particle see Section 2.10. Finally,
as we did for the Lorentz collision model [cf., (2.22), (2.23)], we can write the
dynamical friction and velocity diffusion coefficients for Coulomb collisions of
test particles of species s with a species s& of Maxwellian background particles
in the coordinate-independent vectorial form
"

)v*s/s
s/s"
= S v,
t
"
1 s/s"
)vv*s/s
s/s"
= (v 2 I vv) + $ vv.
t
2

(2.90)
(2.91)

10 For an alternative representation of these various collisional processes using the notation
and functions Chandrasekhar introduced for stellar collisions see Problem 2.27.

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

31

CHAPTER 2. COULOMB COLLISIONS

Figure 2.7: Coulomb collisional effects on the velocity of a test particle: momentum loss or slowing down (S ), angular or perpendicular ( ) and speed
or parallel ($ ) diffusion of the original test particle velocity. The contours
shown are lines of e1 & 0.37 probability (see Section 2.10) at the short times
(t1 , t2 , t3 ) = (0.002, 0.02, 0.2)/ for an energetic electron (me v 2 /2Te = 10) in
an electron-proton plasma for which $ / & 1/40, / & 1/2.

2.8

Test Particle and Plasma Collision Rates+

We now consider the various Coulomb collision effects on typical electrons and
ions in a plasma. For simplicity the plasma will be assumed to be composed of
electrons and one species of ions with charge qi = Zi e, and to have equal electron
and ion temperatures, with both species of particles having Maxwellian velocity
distributions. Thus, the formulas derived in the preceding section will apply.
For illustrative purposes we consider collisional effects on a test electron
and a test ion in the plasma, each having speeds equal to the thermal or most
probable speeds for their respective species:
ve = vT e ,

vi = vT i .

(2.92)

s/s"

for electron-ion (e/i), electronThen, the reference collision frequencies 0


electron (e/e), ion-ion (i/i) and ion-electron (i/e) collisions are simply related:
&
&
e/i
e/e
i/i
e/i
i/e
e/i
0 = Zi 0 , 0 = Zi2 me /mi 0 , 0 = Zi me /mi 0 , (2.93)

in which we have neglected the small differences in ln ss" for differing s and
s& and made use of the quasineutrality condition ne = ni Zi . Further, since
the ratio of ion to electron mass is very large (1836 for protons), we find that
"
the relative speed parameters xs/s defined in (2.82) for the various collisional

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

32

CHAPTER 2. COULOMB COLLISIONS

Figure 2.8: Maxwell integral (x) and related functions.


processes are given by
xe/i =

mi
>> 1,
me

xe/e = 1,

xi/i = 1,

xi/e =

me
<< 1.
mi

(2.94)

Thus, we will need both small and large argument expansions of the Maxwell
integral (x) and its derivative & (x), as well as evaluation of them at the
particular value of unity.
The behavior of (x) and other functions of interest are sketched in Fig. 2.8.
For x = 1, we have = 0.4276, & = 0.4151 and + & /2x = 0.6289. Small
and large argument expansions of interest in evaluating S , , $ , and are:
x << 1

(x) & (4x3/2 /3 )(1 3x/5 + 3x2 /14 ),

& = (2 x ex / ) & (2x1/2 / )(1 x + x2 /2 ),

+ & /2x & (4x1/2 /3 )(1 x/5 + 3x2 /70 ),

(2.95)

(x) & 1 (2 x ex / )(1 + 1/2x 1/4x2 + ),

& (x) = 2 x ex / ,

+ & /2x & 1 1/2x + (ex / x3/2 )(1 1/x + ).

(2.96)

x >> 1

Using only the lowest order of these approximations in (2.87)(2.88), we find


the relationships between various collisional processes listed in Table 2.1. The
e/i
rates are all referred to the electron-ion collision frequency 0 , which is the
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

33

CHAPTER 2. COULOMB COLLISIONS

Table 2.1: Relative Coulomb collision rates for thermal speed test electrons and
ions in a Maxwellian plasma with Te = Ti .

slowing
down
perpendicular
diffusion
speed
diffusion
energy
loss

s/s"

e/i

S / 0

e/i

e/e

0.86/Zi

0.86Zi2

me
mi

0.75Zi

me
mi

1.26/Zi

1.26Zi2

me
mi

1.50Zi

me
mi

0.43/Zi

0.43Zi2

me
mi

0.75Zi

me
mi

0.03/Zi

0.03Zi2

s/s"
e/i
/ 0

s/s"
e/i
$ / 0

me
mi

s/s"

e/i
/ 0

me
2
mi

i/i

e/i

i/e

me
me
0.75Zi
mi
mi
e/i

same as the Lorentz collision frequency in (2.14). The S and components


of the first (e/i) column are the same as those given by the Lorentz collision
model [cf., (2.13) and (2.19)]. All the other electron processes indicated in the
table arise from the finite mass ratio between the electrons and ions, and the fact
that the background particles are in motion. Note that in this general collision
e/i
e/i
model the electron-ion parallel (speed) diffusion ($ ) and energy loss ( )
are of order me /mi << 1 compared to the Lorentz collision model processes
because of the inefficiency of energy transfer in collisions of particles with very
disparate masses.
From Table 2.1 we see that the various
collisional processes naturally split
&
e/i
e/i
e/i
2
into three groups of rates: 0 , Zi me /mi 0 , and Zi (me /mi ) 0 . The
fastest of these rates is the Lorentz collision rate; however, all the electronelectron collisional processes also occur at roughly the same rate and so should
also be taken into account in investigations of electron collisional processes. (The
electron-electron collision processes are small in a plasma where the ions all have
e/i
e/e
e/e
high charge states Zi >> 1 since then 0 = Zi 0 >> 0 .) Physically, on
e/i
this fastest time scale of 1/0 , electron momentum is relaxed by collisions
on both electrons and ions, and the electrons relax within themselves through
all the processes. The electron-electron collisions relax the electrons toward a
Maxwellian distribution (see Chapter
& 11). On the next lower rate or longer
> 43 >> 1 ion-ion collisions
time scale by a factor of order mi /me
relax the ions toward a Maxwellian distribution. Finally, on the longest time
> 1836 >> 1 slower than 1/ e/i , there
scale, which is a factor of about mi /me
0
is energy transfer between the electrons and ions, and ion momentum loss to the
i/e
electrons. [The energy loss rate is negative here because we are evaluating
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

34

CHAPTER 2. COULOMB COLLISIONS

it for a test particle whose energy is below the average particle energy in the
plasma, mv 2 /2 = 3 T /2 see (??)].
Next, we consider the overall plasma electron and ion species collisional
relaxation rates. First, we consider the temperature equilibration rate for a
Maxwellian distribution of test particles of species s colliding with a Maxwellian
s/s"
distribution of background particles s& . Multiplying the defined in the last
line of (2.88) by an isotropic Maxwellian velocity distribution of test particles s
in the form given in (2.76) and using the first property of given in (2.80) to
integrate the contribution to by parts once, we find
3 dTs
s/s"
ns
= ns (Ts Ts" ),
2
dt
where
s/s"

ns

ms
= ns
ms "

(2.97)

,
"
"
4 4 ns ns" qs2 qs2" ln ss"
4
0s/s (vT ss" ) = ns" s /s =
,

{4%0 }2 ms ms" vT3 ss"


(2.98)

is the average energy density exchange rate between the species in which
(2.99)
vT ss" [2(Ts /ms + Ts" /ms" )]1/2 = vT2 s + vT2 s"

is the appropriate mean thermal velocity for a combination of test and background particles, both with Maxwellian distributions. From the equality of
s/s"
s/s"
ns and ns" , it is obvious that
ns

dTs
dTs"
= ns "
,
dt
dt

(2.100)

as required by energy conservation energy lost from the test particle species is
gained by the (dissimilar) background species with which it suffers Coulomb collisions. For a couple of applications of these temperature equilibration formulas
see Problems 2.35 and 2.36.
For the special case of the electron-ion plasma we have been considering, if
we assume Te /me >> Ti /mi (vT e >> vT i ), (2.97) becomes [cf., (2.38)]
me
3 dTe
e/i
ne
= ne (Te Ti ) = 3
ne e (Te Ti ) Qi .
2
dt
mi

(2.101)

In the last expression we have used the e defined in (2.17). The relevant formula
e/i
for the electron-ion energy transfer rate in a plasma with impurities (see
Problem 2.37) was given previously in (2.49) and (2.50).
Finally, we calculate the momentum relaxation rate for two Maxwellian distributions of particles that are drifting (flowing) slowly relative to each other
with velocity V Vs Vs" , assuming |V| << vT ss" . In the rest frame of the
background particles (s& ), the drifting test particle (s) distribution function can
be written as in (2.15):
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

35

CHAPTER 2. COULOMB COLLISIONS


fs (v)

= ns

ms
2Ts
2

$3/2

ns ev /vT s
3/2 vT3 s

&

ms |v V|2
exp
2Ts
,
+
2 v V
+ .
1+ 2
vT s

$
(2.102)

Multiplying the momentum loss rate formula in the first line of (2.88) by this
distribution function and integrating over velocity space, again integrating once
by parts and using the first relation in (2.80), we find
m s ns
where
s/s"
ms ns S

= ms ns

dVs
s/s"
= S ms ns (Vs Vs" ),
dt

(2.103)

,
4 4 ns ns" qs2 qs2" ln ss"
ms s/s"
"
0 (vT ss ) =
3 mss"
3 {4%0 }2 mss" vT3 ss"
(2.104)
4

is the average momentum density exchange rate between the s and s& species of
particles, and vT ss" is the average thermal velocity defined in (2.99). From the
s/s"
symmetric form of ms ns S in terms of the species labels s and s& , it is clear
that the momentum lost from the s species is gained by the s& species and thus
momentum is conserved in the Coulomb collisional interactions between the two
species of particles: ms ns dVs /dt = ms" ns" dVs" /dt.
Specializing again to an electron-ion plasma and assuming as usual that
vT e >> vT i , we find that (2.103) and (2.104) reduce to [cf., (2.34)]
me ne
where
e =

e/i
S

dVe
= me ne e (Ve Vi ) Re ,
dt

4
4 2 ne Zi e4 ln
1
e/i
= 0 (vT e ) =
.
1/2
3/2
e
3
{4%0 }2 3 me Te

(2.105)

(2.106)

This electron momentum relaxation rate is the same as that obtained in (2.17)
for the Lorentz collision model and shows that the fundamental Maxwelliane/i
averaged electron-ion collision frequency e is in fact S . Electron-electron
collisions do not contribute to the momentum relaxation process because they
are momentum conserving for the electron species as a whole. Note also that
the collisional momentum relaxation process acts on the difference between the
electron and ion flow velocities. Thus, the net effect of Coulomb collisions is to
relax the electron flow to the ion flow velocity. Finally, the relevant formula for
e/i
the electron-ion collisional slowing down rate S in a plasma with impurities
(see Problem 2.38) was given in (2.44).
For the slightly fictitious case of two ion species with charge qi = Zi e that
have equal temperatures but are drifting relative to each other with velocity V,
the ion momentum relaxation rate is given by
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

36

CHAPTER 2. COULOMB COLLISIONS


mi ni

dV
= mi ni i V,
dt

where [cf., (2.36)]

4 ni Zi4 e4 ln

i/i

i = S =

1/2 3/2
{4%0 }2 3 mi Ti

(2.107)

1
.
i

(2.108)

i/i

As can be anticipated from the


& S entry in Table 2.1, this momentum relaxation
2
rate is a factor of order Zi me /mi slower than that for electrons. Also, the
i/e
ion-electron collisional effects due to S have been neglected
& in the average ion
momentum loss rate because they are a factor of order me /m
i smaller than
the ion-ion collisional effects. The numerical factor in (2.108) is 2 smaller than
that in (2.106) because of the rest mass and average thermal velocity factors
for this equal mass case. Finally, the relevant formula for the ion-ion collisional
i/i
slowing down rate S in a plasma with impurities (see Problem 2.39) was
given in (2.46).

2.9

Fast Ion Thermalization+

In attempting to heat plasmas one often introduces fast ions (through absorption of energetic neutrals, from radiofrequency wave heating, or directly as
energetic charged fusion products such as particles), which have speeds intermediate between the ion and electron thermal speeds. These fast ions heat
the plasma by transferring their energy to the background plasma electrons and
ions during the Coulomb collisional slowing down process. This fast ion slowing
down and energy transfer process will now be considered in some detail.
For simplicity we consider an electron-hydrogenic (proton, deuteron or triton
mi = 1, 2 or 3 but Zi = 1) background plasma in which both species have
a Maxwellian velocity distribution. The electron and ion temperatures will be
assumed to be unequal, but comparable in magnitude. The fast or test ion
will be allowed to have a mass (mf ) and charge (qf = Zf e) different from
the background ions. Because the fast ion speed is intermediate between the
electron and ion thermal speeds, the relative speed parameters in (2.82) for the
fast ion-ion (f /i) and fast ion-electron (f /e) collisions are given by
xf /i =

mi v 2
v2
= 2 >> 1,
2Ti
vT i

xf /e =

me v 2
v2
= 2 << 1,
2Te
vT e

(2.109)
f /s"

in which v is the fast ion speed. Further, the reference collision frequencies 0
are equal for the electron-hydrogenic ion background plasma:
f /i

f /e

= 0 .

(2.110)

Using the approximations (2.109) in (2.95) and (2.96), we find that the fast
ion transfers energy to the plasma electrons and ions at the rates defined in the
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

37

CHAPTER 2. COULOMB COLLISIONS


last equation in (2.88), which are given to lowest significant order by
f /i

f /e

mf f /i
,
mi 0
mf 4(xf /e )3/2 f /e
mf 4 v 3 f /e

& 2
0 = 2
.
me
me 3 vT3 e 0
3
& 2

(2.111)
(2.112)

f /s"

in (2.87) we see that it depends on v 3 . Thus,


From the definition of 0
f /i
f /e
also depends on v 3 . However, is independent of the fast ion speed
v because the appropriate relative speed for fast ion-electron collisions with
vT e >> v is the electron thermal speed.
Adding together the fast ion energy losses via collisions with background
plasma ions and electrons, the total fast ion energy loss rate becomes
4
5
d
f /e
f /i
= +
,
(2.113)
dt
in which

= mf v2 /2

(2.114)

f /e

is the instantaneous fast ion energy. Since is independent of the fast ion
energy, it is convenient to define a fast ion slowing down time S in terms of it:
S

2
f /e

&

1
f /e
S

1/2

3/2

mf {4%0 }2 3 me Te

=
me (4 2) ne Zf2 e4 ln

mf
me

1
.
e

(2.115)

f /e

Here, the approximate equality to 1/S follows because for mf /me >> 1 the
fast ions are not significantly scattered by the electrons; thus, they lose energy
to the plasma electrons at twice the rate they lose momentum to them. The rate
of transfer of fast ion energy to plasma ions can be referenced to the transfer
rate to the electrons in terms of a critical energy c mf vc2 /2 as follows:
f /i

f /e

4 53/2
c

vc3
,
v3

(2.116)

where

.
/
+ 8
,2/3
mf vc2
3 mf mf
mf
= Te
c
& 14.8 Te
1/3 2/3
2
4
me mi
mp m

(2.117)

in which mp is the proton mass. (For the appropriate modifications when multiple species of ions are present, see Problems 2.40 and 2.50.) In terms of this
critical energy, (2.113) can be written as
+
4 53/2 ,
2
d
c
=
1+
.
(2.118)
dt
S

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

38

CHAPTER 2. COULOMB COLLISIONS

Figure 2.9: Fast ion energy transfer rate versus energy . The energy transfer
is primarily to electrons for > c , but to ions for < c .
The fast ion energy transfer rates as a function of energy are illustrated in Fig.
2.9. For fast ion energies greater than c the energy transfer is primarily to
electrons, while for < c it is primarily to ions.
Since (2.118) applies for all fast ion speeds between the electron and ion
thermal speeds, it will be valid for all fast ion energies during the thermalization
process. Thus, its solution will give the fast ion energy as a function of time as it
transfers its energy to the background plasma. To solve (2.118)
it is convenient
&
to convert it to an equation for the fast ion speed v 2/mf , for which it
becomes
+
,
v
dv
vc3
=
1+ 3 ,
(2.119)
dt
S
v
vc

+
,1/3
2 c
3 me
=
vT e .
mf
4 mi

(2.120)

Multiplying (2.119) by v 2 and integrating over time from t = 0 where the initial
fast ion speed will be taken to be v0 to the current time t where it has speed v
(assumed > vTi ), we obtain
# 3
$
v0 + vc3
S
ln
t=
,
(2.121)
3
v 3 + vc3
or

v 3 (t) = (v03 + vc3 ) e3t/S vc3 .

(2.122)

The fast ion energy (t) = mf v 2 (t)/2 during the slowing down process can be
readily obtained from this last result.
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

39

CHAPTER 2. COULOMB COLLISIONS


10

E/Ec
3
1
.3
.1
0

0.5

t/s
Figure 2.10: Decay of fast ion energy versus time during thermalization into
a background plasma for various initial ratios of to the critical energy c .

The decay of the fast ion energy with time is illustrated in Fig. 2.10. Note
that for initial energies much greater than the critical energy c the fast ion
f /e
energy decays exponentially in time at a rate 2/S = due to collisions with
electrons, as is apparent from (2.118). However, when the energy drops below
c the energy transfer is predominantly to the ions and the fast ion energy
decays much faster than exponentially. The total lifetime for thermalization (to
v & vT i << vc ) of the fast ion into the background plasma ions is
f & (S /3) ln [1 + (0 /c )3/2 ] = (S /3) ln [1 + v03 /vc3 ].

(2.123)

A couple of applications of these fast ion slowing down effects and formulas are
developed in Problems 2.412.42.
Next, we calculate the fraction of the fast ion energy that is transferred to
the background plasma electrons and ions over the entire fast ion slowing down
process. Since in many plasma situations the fast ions are also susceptible to
other, direct loss processes such as charge-exchange, we introduce a probability
exp(t/cx ) that the fast ion will remain in the plasma for a time t against
charge-exchange losses at rate 1/cx . Then, the fraction Ge of the total fast
ion energy 0 mf v02 /2 transferred to the electrons during the thermalization
process is given by
$ f /e
+
, /3
% f #
% v0 3
1
d et/cx
2
v v dv v 3 + vc3 S cx
dt
= 2
.
(2.124)
Ge
0 0
dt f /e + f /i
v0 0 v 3 + vc3 v03 + vc3
Similarly, the fraction Gi of fast ion energy transferred to the ions is (for the
simpler case where cx , see also the form given in Problem 2.43)
$ f /i
+
, /3
% f #
% v0 3
1
d et/cx
2
vc v dv v 3 + vc3 S cx
dt
= 2
.
(2.125)
Gi
0 0
dt f /e + f /i
v0 0 v 3 + vc3 v03 + vc3
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

40

CHAPTER 2. COULOMB COLLISIONS


1

Gi,!Ge

0.8
ions
0.6

0
0.1

= 0
1

3
10

0.4
0.2

s
____
cx

0
1

3
10

electrons
0.2

0.5

Eo/Ec

10

Figure 2.11: Fraction of the fast ion energy 0 that is transferred to the background plasma electrons (Ge ) and ions (Gi ) as a function of the ratio of the
initial energy 0 to the critical energy c . The variation with the parameter
S /cx indicates the influence of direct fast ion losses (at rate 1/cx ) during the
thermalization process.

The fraction of fast ion energy lost due to charge-exchange is 1 Gi Ge .


However, a portion of this energy may be absorbed in the plasma if some of the
fast neutrals produced by charge-exchange are reabsorbed before they leave the
plasma.
The variation of the fractions Ge , Gi of fast ion energy transferred to plasma
electrons and ions during the thermalization process are illustrated in Fig. 2.11.
< 2.5 c , which is above
Note that the integrated fractions become equal for 0
the value of 0 & c where the instantaneous energy transfer rates are equal.
> 1, and can greatly
Also, charge-exchange losses become significant for S /cx
diminish the fast ion energy transfer to the plasma for S /cx >> 1. For some
typical applications of fast ion slowing down and energy transfer processes and
their effects on plasmas, see Problems 2.442.47.
In addition to energy loss, the fast ions experience perpendicular and parallel
diffusion in velocity space during their thermalization. The relative importance
of the various Coulomb collision processes on the fast ion for the conditions
given in (2.109) are indicated in Table 2.2. From this table we see that for
>> c the momentum and energy losses by the fast ions to the electrons
are the dominant processes because then the velocity space diffusion effects
indicated by , $ are small. However, for < c the fast ions lose energy
primarily to the background ions and their perpendicular or angular diffusion
rate in velocity space becomes equal to their energy loss rate. For some typical
applications of fast ion scattering processes and their effects on plasmas, see
Problems 2.482.50.
The energy or speed diffusion process indicated by $ is negligible until
the fast ion energy is reduced to approximately the ion temperature in the
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

41

CHAPTER 2. COULOMB COLLISIONS

Table 2.2: Relative Coulomb collision rates for fast ions with vT i << v << vT e
slowing down in a plasma with Maxwellian electrons and ions (c 14.8 Te ).
f /i
f /e
slowing
down

f /s"

f /e

S / 0

perpendicular
diffusion
speed
diffusion
energy
loss

1+

mf
mi

f /s"
f /e
/ 0

f /s"
f /e
$ / 0

Ti
<< 1
mf v 2 /2

f /s"

f /e

/ 0

mf
mi

mf 4

me 3

vT e

3
mf 8

me 3

v
vT e

$3

v
vT e
v
vT e
$3

<< 1

<< 1

=2

$3/2

$3/2

background plasma. Since the energy diffusion process is thus negligible during
the fast ion thermalization process, and the perpendicular diffusion has no effect
on the energy transfer rates, our characterization of the fast ion slowing down
process as one of a monotonic decrease in the fast ion energy is a reasonably
accurate one. A kinetic description that allows for pitch-angle () scattering
along with the fast ion energy slowing down is developed in Section 11.4.

2.10

Evolution of Velocity of a Test Particle+

To further illustrate the Coulomb collision effects, we examine the collisional


evolution of the velocity of a test particle for short times where the velocity
changes are small. The test particle will be assumed to be colliding with a
plasma whose components have Maxwellian distributions. Thus, the results of
Section 2.7 will be applicable.
A test particle of species s will be taken to have an initial velocity v0 in the

ez or parallel direction (cf., Fig. 2.4). Integrating the first equation of (2.88)
over a short time t >> t (for validity of the dynamical friction and velocity
diffusion coefficients), we find that the mean parallel or
ez component of the
test particle velocity after a time t is
2 s/s"
v$ = v0 (1 S t), S = Ss
S (v0 ).
(2.126)
s"

This result is valid for S t << 1 and indicates the monotonic decrease in test
particle momentum due to Coulomb collisons. Similarly, the test particle energy
after a short time t can be obtained directly by integrating the last equation of

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

42

CHAPTER 2. COULOMB COLLISIONS


(2.88) over time:
1
1
mv 2 = mv02 (1 t),
2
2

= s

s/s"

(v0 ).

(2.127)

s"

Thus, the average test particle speed v, which will be used below, will be defined
by
&
v v 2 & v0 (1 t/2).
(2.128)

Similar to (2.126), these formulas are only valid for t << 1.


The angular (perpendicular) velocity and speed (parallel) diffusion processes
in velocity space indicated by and $ have to be treated differently. Because
these Coulomb collision effects are random in character and diffusive, they lead
to a Gaussian probability distribution P (v v0 ) of the velocity about the initial
test particle velocity v0 . Since the diffusion results from purely random processes, we can anticipate (and will derive in Chapter 11) that this probability
distribution will be Gaussian and of the form:
.
/.
/
2
2
2
2
1
e /2
e(vv) /2#

P (v v0 ) =
.
(2.129)
2

2 v 2
2 $

Here, 2 [arcsin1 (v /v)]2 & (vx2 + vy2 )/v 2 and v is as was defined in (2.128).
Taking velocity-space
' averages 'of various quantities A(v) over this probability
distribution [ A 0 2v 2 dv 0 sin d P (v v0 ) A(v)], we find that while
the average of the diffusive deflections vanish [v x = v y = 0, v$ v $ = 0], the
diffusive spreads in the perpendicular and parallel (to v0 ) directions are
2 = v2 + v2 = 2 2 v2 ,
v 2 v02 & v
x
y
0

(v$ v0 )2 = (v$ v$ )2 = $2 .

(2.130)

To determine the probability variances and $ for the diffusive Coulomb


collisional processes, we integrate the middle two equations in (2.88) over a short
time t, and obtain
2 s/s"
s
2 = ( t) v 2 ,
v
=

(v0 )
(2.131)

0
s"

(v$ v$ )2 = ($ t) v02 ,

$ = $s

2
s"

s/s"

$ (v0 ).

(2.132)

Comparing (2.130), (2.132) and (2.132), we see that for Coulomb collisions
&
&
(2.133)
= t/2, $ = $ t v0 .

The relative collisional spreads (half-widths in velocity space to points where


the probability distribution drops to e1/2 & 0.61 of its peak value) of the test
particle velocity in the directions perpendicular and parallel (i.e., for speed or
energy diffusion) relative to its initial velocity v0 are given by
&
&
(2.134)
& v /v0 = t/2, v/v0 $ /v0 = $ t.

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

43

CHAPTER 2. COULOMB COLLISIONS

Note that
& in comparing the perpendicular diffusion factor in (2.129) for which
= t/2 with the perpendicular diffusion in the Lorentz model as
given
in (2.21), we need to realize that = 2 and hence that = t for
the Lorentz collision model. These formulas indicate that, even for very high
temperature plasmas with n3D >> 1 where the Coulomb collision rates are very
slow, only a short time is required to diffuse the test particle velocity through
a small v << v0 . For example, as indicated in (2.20), the time required to
diffuse a particles velocity through a small angle & v /v0 << 1 is only
t & 2 2 / << 1/ [or, t & 2 / << 1/ for the Lorentz collision model].
Thus, because of the diffusive nature of Coulomb collisions, it takes much less
time to scatter through an angle << 1 in velocity space than it does to scatter
through 90 ( 1). The various diffusive collisional effects are illustrated in
Fig. 2.7. There, the contours shown indicate where the probability distribution
P in (2.129) is equal to e1 & 0.37 of its peak value for t = 0.002, 0.02, 0.2,
for a typical set of test particle parameters.
The change in the average energy mv 2 /2 can also be obtained using (2.126),
(2.130) and (2.133). This procedure yields, correct to first order in t << 1,
6
7
6
7
1
2 = 1 m v 2 + v 2 = 1 m v 2 + 2(v v )v + (v v )2 + v 2
mv
$
$
$
$
$
$

$
2
2
2
&

1
2
2 m[v0 (1

2
2S t) + $2 + 2
] = 12 mv02 [1 (2S $ )t].

This result is the same as (2.127) because of the relation between the various
collisional processes given in (2.89).
The formulas developed in this section also provide a basis for a probablistic
(Monte Carlo) numerical approach for inclusion of Coulomb collision effects in
other plasma processes such as single particle trajectories. Thus far we have
found that after a short time t a test particles velocity and speed decrease
according to (2.126) and (2.128). However, the test particle also acquires a
diffusive spread in the perpendicular and parallel directions as given by (2.129)
with the spreads (variances) defined in (2.133). Further, the velocity space
latitudinal angle [cf., (2.1)] is completely randomized by successive individual
Coulomb collisions in a time scale t >> t. Hence, defining a random
variable to be evenly distributed between 0 and 1, and independent random
variables 1 , 2 sampled from a normal probability distribution [i.e., Gaussian
such as indicated in the $ part of (2.129)] with zero mean and a mean square
of unity (i.e., 1 = 2 = 0 but 12 = 22 = 1), we find that the total velocity
vector v after a short time t (t << 1) can be written as
6
7
&
&
ez ( 1 + 1 $ t ) + |2 | t/2 (
ex cos 2 +
ey sin 2) .
v = v0 (1 t/2)
(2.135)

This form can be used to develop a Monte Carlo algorithm for advancing
the test particle velocity v taking into account the Coulomb collision dynamical friction and velocity space diffusion effects occurring over the short time
t << 1/. Since (2.135) implies a change in the velocity of the test particle,
DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

44

CHAPTER 2. COULOMB COLLISIONS

in order to preserve the momentum and energy conserving properties of the


elastic Coulomb collision process, the velocity of the background particles must
also change, at least on average. Hence, in order to develop a complete Monte
Carlo-based Coulomb collision operator we should consider simultaneously both
a test and a background particle. Then, the change in velocity v v v0
for the test particle is determined from (2.135), and that for the background
particle is given by v& = (ms /ms" )v see (2.59).
REFERENCES AND SUGGESTED READING
The basic Coulomb collision processes were first worked out in the analogous context (see Problem 2.9) of the gravitational interaction of stars:
Chandrasekhar, Principles of Stellar Dynamics (1942).
S. Chandrasekhar, Rev. Mod. Phys. 15, 1 (1943).
A comprehensive application to Coulomb collisions in a plasma was first presented in
Spitzer, Physics of Fully Ionized Gases (1962).
The most general development of the dynamical friction and velocity diffusion coefficients for a plasma in terms of the Rosenbluth potentials originated in the paper
M.N. Rosenbluth, W. MacDonald and D. Judd, Phys. Rev. 107, 1 (1957).
The most comprehensive treatments of the Coulomb collision effects on test particles
in a plasma are found in
B.A. Trubnikov, Particle Interactions in a Fully Ionized Plasma, in Reviews of
Plasma Physics, M.A. Leontovich, ed. (Consultants Bureau, New York, 1965),
Vol. I, p. 105.
D.V. Sivukhin, Coulomb Collisions in a Fully Ionized Plasma, in Reviews of
Plasma Physics, M.A. Leontovich, ed. (Consultants Bureau, New York, 1966),
Vol. IV, p. 93.
A brief, but very useful summary of the important Coulomb collision formulas in this
chapter is given in
Book, NRL Plasma Formulary (1990), p. 31.
A book devoted almost entirely to the subject of Coulomb collision effects in a plasma
with numerous examples worked out is
Shkarofsky, Johnston and Bachynski, The Particle Kinetics of Plasmas (1966).
Also, most books on plasma physics have chapters devoted to discussions of Coulomb
collision effects. Among the most descriptive and useful are those in
Spitzer, Physics of Fully Ionized Gases (1962), Chapter 5
Rose and Clark, Plasmas and Controlled Fusion (1961), Chapter 8.
Schmidt, Physics of High Temperature Plasmas (1979), Chapter 11.
Krall and Trivelpiece, Principles of Plasma Physics (1973), Chapter 6.
Golant, Zhilinsky and Sakharov, Fundamentals of Plasma Physics (1980), Chapter 2.
The original theory of runaway electrons was developed in
H. Dreicer, Proceedings of the Second United Nations International Conference
on the Peaceful Use of Atomic Energy (United Nations, Geneva, 1958), Vol. 31,
p. 57. See also, Phys. Rev. 115, 238 (1959).

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

45

The thermalization of a fast ion in a Maxwellian plasma was first developed in


D.J. Sigmar and G. Joyce, Nuclear Fusion 11, 447 (1971).
T.H. Stix, Plasma Physics 14, 367 (1972).
Inclusion of charge-exchange loss and geometry effects on neutral-beam-injected fast
ions are discussed in
J.D. Callen, R.J. Colchin, R.H. Fowler, D.G. McAlees and J.A. Rome, Neutral
Beam Injection into Tokamaks, Plasma Physics and Controlled Nuclear Fusion
Research 1974 (IAEA, Vienna, 1975), Vol. I, p. 645.
The Monte Carlo computational approach to including Coulomb collisional effects has
been developed primarily in the context of investigating transport processes in
R. Shanny, J.M. Dawson and J.M. Greene, Phys. Fluids 10, 1281 (1967).
K.T. Tsang, Y. Matsuda and H. Okuda, Phys. Fluids 18, 1282 (1975).
T. Takizuka and H. Abe, J. Comput. Phys. 25, 205 (1977).
A.H. Boozer and G. Kuo-Petravic, Phys. Fluids 24, 851 (1981).

PROBLEMS
2.1 Consider the length scales relevant for electron Coulomb collision processes in a
typical university-scale magnetic fusion plasma experiment that has ne = 21013
cm3 and Te = Ti = 1 keV. Calculate: a) the distance of closest approach bmin ;
b) the average interparticle spacing; c) the maximum interaction distance bmax ;
and d) the average collision length e = vT e /e for electrons in this plasma.
What is the ratio of each of these lengths to the mean interparticle spacing? /
2.2 Consider the length scales relevant for electron Coulomb collision processes in
a laser-produced electron-proton plasma that has ne = 1029 m3 and Te = Ti
= 1 keV. Calculate: a) the distance of closest approach bmin ; b) the average
interparticle spacing; c) the maximum interaction distance bmax ; and d) the
average collision length e = vT e /e for electrons in this plasma. What is the
ratio of each of these lengths to the mean interparticle spacing? /
2.3 Estimate the time scales relevant for electron Coulomb collision processes in the
earths ionosphere at a point where ne = 1012 m3 , Te = 1 eV. For simplicity,
use a Lorentz collision model and assume the ions have Zi = 1 and Ti " Te .
Calculate the times for: a) a typical Coulomb interaction at the average interparticle spacing; b) an electron to traverse the Debye shielding cloud; and c) the
average electron collision time e = 1/e . How long (or short) are each of these
1
times compared to the plasma period pe
? /
2.4 a) Show that for a Lorentz collision model an electron with an initial velocity v0
loses momentum exponentially in time at a decay rate given by (v0 ). b) What
is the electron energy after its momentum is totally depleted? c) Calculate the
distance the electron travels in its original direction of motion while losing its
momentum. d) Evaluate the momentum decay rate and distance the electron
travels for a plasma with Zi = 5, ne = 1019 m3 , Te = Ti = 100eV , and an
initial electron test particle energy of 1 keV. [Hint: Be careful to distinguish
between an electrons velocity (a directional, vector quantity) and its speed (a
scalar quantity).] //
2.5 In Section 2.1 we derived the momentum impulse v for a single Coulomb
collision in the Lorentz collision model by integrating me (dv/dt) = qe E over an

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

46

CHAPTER 2. COULOMB COLLISIONS

unperturbed straight-line trajectory to determine v and then used v# "


v v /2v. Show that the result for v# given in (2.6) can be obtained

directly by integrating E[x(t)] along a perturbed electron trajectory x(t) = x+ x


due to the E field of the ion on the electron
that includes the first order effects x
, and then make use
trajectory. [Hint: First calculate the perturbed velocity v
of the fact that
(
,
+
%
%
qe
qe
E ((
) =

v# =
ez
dt E(x + x
dt
ez E(x) + x
+

me
me
x (x
%
%
% t
t
e
Zi e2
dt% Ez (t% ) =
=
dt v
dt v# 2
.] ///
2 t2 )3/2
me
{4'
}m
v
(b
+
v
0
e

2.6 Show that the Lorentz collision frequency can be derived from the Rutherford
differential scattering cross-section d = (d/d)d given in (??) and Problem 2.24. First, show that for the Lorentz collision model the scattering angle
is given for typical small-angle Coulomb collisions by " 2Zi e2 /({4'0 }mv 2 b)
= 2bcl
min /b, and that the differential scattering cross-section kernal is d/d
2
4
= |(b db d)/(d d cos )| " (b/)|db/d| = 4(bcl
min ) / . Then, determine the
effective cross-section for momentum transfer m , which is defined by
%
m d (d/d) (1 cos ).
In performing this integral discuss the maximum, minimum scattering angles
max , min in terms of the bmin , bmax interaction distances. Finally, show that
= ni m v yields the Lorentz model collision frequency given in (2.14). ///
2.7 Use the full Rutherford differential scattering cross-section and the procedure
outlined in the preceding problem to give an alternate derivation of the Lorenz
collision frequency that takes into account classical hard, or large angle collisions; i.e., do not initially assume << 1. ///
2.8 The Lorentz collision frequency can also be determined from the Langevin
equation
me

dv
= me v + F(t)
dt

in which me v is the dynamical friction force, and F(t) is a stochastic force,


which for Coulomb collisions is that given by (2.2). Assuming is constant in
time, use an integrating factor et in solving the Langevin equation to determine
the particle velocity v(t) after its initialization to v0 at t = 0. Next, calculate
the ensemble average of the electron kinetic energy as a function of time, and
show that it yields
%
%F(0) F( )&
1 e2t
%v 2 (t)& " v02 e2t +
d
.
2
m2e

Show that for times long compared to the duration of individual Coulomb collisions but short compared to the momentum loss collision time (b/v << t <<
1/), the electron kinetic energy is constant through terms of order t when
%
1

d %F(0) F( )&.
2m2e v 2

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

47

Thus, is proportional to the autocorrelation function of the Coulomb collision force F. Next, show that for Coulomb collisions between electrons and
a stationary background of randomly distributed ions this formula yields the
Lorentz collision frequency given by (2.14). Finally, making use of the equilibrium (t ) statistical mechanics (thermodynamics) property that for
random (Brownian) motion due to a stochastic force F the ensemble-average
kinetic energy me %v 2 &/2 of a particle is Te /2, note that this last result yields
%
%F(0) F( )&
me
=
d
, fluctuation-dissipation theorem,
2 Te
m2e
which is also related
for
' to Nyquists
' theorem
'
' 2noise in electric circuits. [Hints:
a) Here, %f & ni d3x f = ni dz b db 0 d f ; b) The electron position
for an ion at x = z
ez at time t = 0 is x = b(
ex cos +
ey sin ) (z vt)
ez . ///

2.9 Consider cumulative small-angle collisional interactions of a test star of mass Mt


and velocity v in a galaxy for which the gravitational force between it and groups
of field stars with density nf and mass Mf is given by [cf., (??) in Section A.6]
FG = GMt Mf (xt xf )/|xt xf |3 an attracive inverse square law force like
that for Coulomb collisions of oppositely charged particles. Develop a model
for collisions of this test star with other, background stars and show that the
reference gravitational collision frequency analogous to (2.14) is
+
,
4G2 nf Mf2
D0 u2
G =
ln
,
v3
G(Mt + Mf )
1/3

where bmax has been taken to be D0 , the mean distance between stars ( nf
)
and bmin has been taken to be the minimum interaction distance given by an
expression analogous to that implied by (2.9). Estimate the time in (years) for
the velocity of our sun to scatter through 90 , assuming that our sun is a typical
star in our galaxy which has a mass of 2 1030 kg, a velocity of 20 km/s and
a mean separation from other stars in our galaxy of 1 parsec (" 3 1013 km).
Will you be concerned about this scattering process in your lifetime? ///
2.10 Estimate the diffusion coefficient D for a Ficks law representation ( = Dn)
of the particle flux due to electron-ion Coulomb collision effects in an inhomogeneous plasma as follows. For simplicity, use a Lorentz collision model
and assume the electrons have a density gradient but no temperature gradient.
Then, show by balancing the pressure gradient force density pe = Te ne
against the frictional drag induced by Coulomb collisions that D = Te /me e
= e 2e /2. Estimate the magnitude of this diffusion coefficient for the plasma
described in Problem 2.3 and compare the result to the viscous diffusion coefficient for molecules of air at the earths surface. Why is the diffusivity of charged
particles in a plasma so much larger? //
2.11 Estimate the D-D fusion reaction rate (use f v " 1017 cm3/s) for a plasma
with Te = Ti = 40 keV, and ne = 1020 m3 . Compare this rate to typical
electron and ion Coulomb collision rates in this fusion plasma. How many times
do electrons and ions scatter through 90 during a typical D-D fusion in this
plasma? How far do typical electrons and ions travel in a characteristic fusion
reaction time? /
2.12 Consider the angular scattering of a beam of 100 eV electrons introduced into
an Argon laboratory plasma that has Te = 3 eV, Ti = 1 eV, Zi = 3 and an

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

48

CHAPTER 2. COULOMB COLLISIONS

electron density ne = 1019 m3 . Using the Lorentz collision model, estimate


the distance over which the beam electrons are: a) scattered through an angle
of 6 ; b) scattered via small angle collisions through an angle of about 90 ; and
c) deflected 90 via hard collisions. Finally, d) estimate the angle through
which a beam electron is scattered in a typical Coulomb interaction when the
impact parameter b is given by the mean interparticle spacing of ions. //
2.13 Determine the energy at which electrons run away in response to an electric
field in an impure plasma as follows. Assume that a nearly Maxwellian plasma
(Ti Te ) is composed of electrons and various
0 species of ions with charge Zi ,
for which charge neutrality requires ne =
i ni Zi . Calculate the frictional
drag force on electrons in the high energy tail (me v 2 /2 >> Te ) of the electron
distribution. Show that the energy at which electrons run away is given by
(2.27) with Zi now replaced by the Zeff defined in (2.43). Also, estimate the
fraction of electrons that are runaways for |E|/ED = 0.1 and Zeff = 2. //
2.14 As electrons become relativistic (v c) the dynamical friction force decreases
less rapidly than the 1/v 2 indicated in (2.25) and Fig. 2.6. In fact, it becomes
nearly constant for (1 v 2 /c2 )1/2 >> 1. Then, if the electric field is
weak enough, there are no runaway electrons. Determine the dynamical friction
force on relativistic electrons in a nonrelativistic plasma composed of electrons
and ions of charge Zi as follows. First, show that the change in perpendicular
momentum (p me v) in a single Coulomb collision is given by
p =

2Zi e2
(
ex cos +
ey sin ).
{4'0 }bv

&
Next, use the relativistic form of the total particle energy ( = m2e c4 + p2 c2 )
to show that for Coulomb scattering (constant energy) collisions between high
energy electrons and background electrons or ions of mass mi the change in
parallel momentum is
+
,
p p
me
p# "
1+
.
2p
mi
Show that the frictional force induced by Coulomb collisions of the high energy
electron with the background plasma is thus
#
$
4 ne e4 ln
1 + Zi
%F# & "
1+
{4'0 }2 me v 2

Finally, show that for a weak electric field satisfying


|E|/ED < 2Te /(me c2 )
no runaway electrons will be produced in the plasma. ///+
2.15 Estimate the electric field strength at which the entire electron distribution
function runs away as follows. First, assume the ions are at rest and the
electrons are described by a shifted Maxwellian as defined in (2.102). Then,
transform to the electron rest frame where V = 0. In this frame the ions all
have a velocity V. Show that the frictional force on a test ion is given by
i/e
(mi /me ) mi 0 (v) i/e (x) V in which xi/e = V 2 /vT2 e . Find the maximum of
this frictional force as a function of V /vT e (cf., Fig. 2.8). Then, use the fact

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

49

that this frictional force must be equal and opposite to the maximum force on
the electron distribution to estimate the critical electric field strength (in terms
of the Dreicer field) for total electron runaway. Also, show that at this electric
field strength an average electron is accelerated to roughly its thermal speed in
an appropriate electron collision time. //+
2.16 At what electron temperature is the electrical resistivity of an electron-proton
plasma with ln 17 the same as that of copper at room temperature for which
" 1.7 108 m? /

2.17 In a typical university-scale tokamak experiment an electron-proton plasma with


ln 17 is heated to a temperature of about 300 eV by the joule or ohmic
heating induced by an electric field of about 0.5 V/m. What current density
(in A/cm2 ) does this electric field induce in such a plasma? What is the joule
heating rate (in W/cm3 )? /
2.18 Determine the plasma electrical impedance to an oscillating electric field as follows. First, assume a sinusoidal electric field oscillating at a (radian) frequency
eit . Then, solve an appropriate electron fluid momentum density
: E(t) = E
equation and show that the frequency-dependent electrical conductivity can be
written as
ne e2

() =
.
me (e i)
Over what frequency range is the plasma resistive (dissipative) and over what
range is it reactive? What frequency ranges (in Hz) are these in the earths
ionosphere for the parameters of Problem 2.3? //
2.19 The plasma electrical conductivity is modified in a plasma with neutral particles. Add a neutral friction force me ne en Ve , where en = nn en v is the
Maxwellian-averaged electron-neutral collision frequency, to the right of (2.29)
and show that in equilibrium the modified electrical resistivity is given by
me (e + en )
=
. //
ne e2
2.20 Determine the neutral density range over which the effects of neutral particles
on the electrical conductivity can be neglected using the result given in the
preceding problem as follows. The reaction-rate en v for ionization of atomic
hydrogen by electrons is given approximately (to within about a factor of two)
by
en v " 1.5 108 cm3/s for 10 eV Te 104 eV.
How small must the ratio of the neutral to electron density (nn /ne ) be to neglect electron-neutral collision effects in an electron-proton plasma for Te =
10, 102 , 103 , and 104 eV? Explain why this density ratio varies so dramatically
with electron temperature. /
2.21 In high neutral pressure, low temperature, partially ionized plasmas (e.g., in the
glow discharge in fluorescent light bulbs), electron-neutral collisions compete
with Coulomb collisions. In particular, they can become dominant in the high
energy tail of the electron distribution function, thereby causing it to effectively
vanish for energies above a cut-off energy. Estimate the cut-off energy for
a Te = 3 eV, ne = 1010 cm3 electron-proton plasma that has a hydrogen
neutral density determined by a 3 mm Hg filling pressure, assuming an electronionization rate coefficient en v = 1010 cm3/s for this Te = 3 eV plasma. /

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

50

CHAPTER 2. COULOMB COLLISIONS

2.22 Sketch the variation of the energy transfer rate Qi in (2.38) from electrons to
ions in a Maxwellian electron-proton plasma as a function of Te /Ti . Find the
value of Te /Ti at which the maximum energy transfer occurs. Explain physically
why the energy transfer rate decreases for increasing Te /Ti >> 1. /
2.23 Consider the thermal equilibration of an electron-ion plasma with Te > Ti .
Eliminating Ti in favor of the final temperature T = (Te + Ti )/2, show that in
the absence of joule heating, Eq. (2.41), which governs the electron temperature
evolution, can be reduced to
dz
z1
=
,
dt
z 3/2

1/2

3/2

m i e 1
mi {4'0 }2 3 me T

=
me 4 z 3/2
me 16 2 ne Zi e4 ln

in which z Te /T . Integrate this equation to obtain in general


( 1/2
(
(z
t
1 (( 2 3/2

= ln (( 1/2
+ z
+ 2 z 1/2 + C,

z
+ 1( 3

where C is a constant to be determined from the initial conditions. Estimate the


range over which Te decays exponentially in time toward T and indicate the
decay rate. Discuss the relationship of this decay rate to a simple one derivable
from (2.39) with Te fixed and e = constant. //
2.24 Utilize the Rutherford differential scattering cross-section
d
qs2 qs2"
1
=
,
d
4u4 m2ss" sin4 /2

tan

qs qs"
bcl
=
= min ,
2
2
mss" u b
b

in which is the scattering angle, to give an alternate derivation of the frictional


"
"
drag and velocity diffusion coefficients %v&s/s/t and %vv&s/s/t defined
in (2.60), (2.61) that takes into account classical hard, or large angle collisions.
Show that the results obtained this way reduce to those given in (2.65) and (2.66)
in the limit ln >> 1. [Hint: bdb d = d = (d/d) d = (d/d) sin d d
and after the collision the test particle velocity in the center-of-momentum frame
is given by u + u = (
ex sin cos +
ey sin sin +
ez cos ) u.] ///+
2.25 Show that the rate of momentum and energy loss of a test particle of species s
by collisions with background particles having an arbitrary velocity distribution
fs" (v) can be written, in analogy with electrostatics, as
#
$
#
$
dv

d ms v 2

mss"
ms
= Qs
,
= Qs v

,
dt
v
dt
2
v
ms
where the analogous potential and charge Qs are defined by
%
f " (v% )
ms
(v) Hs" (v) =
d3v % s
, Qs ms ss" .
mss"
|v v% |
Show that for an infinitely massive, immobile background (Lorentz collision
model) these formulas reduce, in analogy with an electrostatic point charge Qs
at the origin of velocity space, to
#
$
dv
n"
d ms v 2
ms
= Qs s3 v = ms (v)v,
= 0. ///+
dt
v
dt
2

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

51

CHAPTER 2. COULOMB COLLISIONS

2.26 Use the formulas derived in the preceding problem to consider collisions of a
test particle s with a spherically symmetric velocity distribution of background
particles that all have the same speed V : fs" (v) = (ns /4V 2 )(v V ). Show
that for these collisional processes the analogous potential is given by
:
ms
1/V, v < V,
=
ns "
1/v, v > V.
mss"
Calculate the momentum and energy loss rates for test particle (s) speeds v <
V , and v > V . Discuss the results obtained in analogy with electrostatics,
and in particular explain by analogy with electrostatics why the test particle s
exchanges no momentum with the background when v < V. ///+
2.27 In the original work on stellar collisions Chandrasekhar introduced the function
% z
2
(z) z% (z)
2

G(z)
,
where
(z)

dy ey = erf (z).
2z 2
0
Show that this Chandrasekhar function G is related to the Maxwell integral
by

(x) vT2 s"

G( x ) =
=
,
2v 2
2x
and hence that the various collision frequencies for Coulomb collisions of a test
particle of species s with a Maxwellian distribution of background particles of
species s% can be written as
#
$#
$
2Ts
m " G(v/vT s" )
s/s"
S
= ss"
1+ s
,
Ts"
ms
(v/vT s )
s/s"

s/s

"

=
=

2 ss" [(v/vT s" ) G(v/vT s" )] (vT3 s /v 3 ),


2 ss" G(v/vT s" )(vT3 s /v 3 )

in which
s/s"

ss" 0

(v)

v3
4 ns" qs2 qs2" ln ss"
=
vT3 s
m2s vT3 s

is the reference collision frequency, which has the advantage of being independent
of the particle speed v. //+
2.28 Discuss the changes that occur in Problem 2.4 when general Coulomb collisions
are allowed for instead of the Lorentz collision model. In particular, indicate the
approximate magnitude and direction of changes in the momentum decay rate,
the distance traveled and the rate of energy transfer to the ions for the plasma
parameters indicated. //+
2.29 Consider the Coulomb collision scattering processes on a D T fusion-produced
particle ( = 3.52 MeV) in a thermonuclear plasma (50% D, 50% T , Te = Ti
= 10 keV, ne = 1020 m3 ). Calculate the collision rates for slowing down (S ),
perpendicular diffusion ( ), parallel diffusion (# ) and energy loss ( ) of the
particle in the plasma. Discuss which of the collisional processes (/e, /D,
or /T ) dominate each of these rates and why. How long will it take such a
fusion-produced particle to deposit half of its energy in the plasma? /+

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

52

CHAPTER 2. COULOMB COLLISIONS

2.30 The direction of energy transfer in Coulomb collisions of test particles with
a background plasma depends on the test particle energy and other parameters. Estimate the particular test particle energies at which there is no energy
exchange between test electrons, test protons and a Maxwellian background
electron-proton plasma that has Te ,= Ti , but Te of the same order of magnitude
as Ti . /+
2.31 Evaluate the ratio of energy diffusion to energy loss for electrons on the high
energy tail (me v 2 /2 > Te ) of a Maxwellian electron distribution function. Use
this result to: a) find the probability that a tail electron will gain rather than
lose energy; and b) discuss phenomenologically how the energy dependence of
the Maxwellian tail of the electron distribution function is determined. //+
2.32 Consider a test electron with an energy of 10% of the electron temperature in a
plasma. Show that the electron gains energy approximately linearly with time
from a Maxwellian background of electrons. Estimate the time required (in
terms of e ) for the test electron to acquire an energy approximately equal to
the plasma electron temperature. //+
2.33 It is of interest to drive the current in a tokamak plasma by means other than
via the usual inductive electric field. Thus, one often seeks [see N.J. Fisch,
Rev. Mod. Phys. 59, 175 (1987)] to drive currents by radiofrequency waves
that impart momentum to a selected group of suprathermal (v >> vT e ) electrons. Coulomb collision effects relax these suprathermal electrons back into
the background distribution and thus limit the current produced. Estimate
the steady-state efficiency J/Pd for such a process as follows. Consider a
suprathermal electron that has a large velocity v0 relative to the thermal speed
of background electrons which will be assumed to have a Maxwellian velocity
distribution. Assume the ions in the plasma have charge Zi , a Maxwellian distribution, and a comparable temperature to the background electrons. Show that
the z-directed velocity component and speed of the suprathermal electron are
e/e
e/e
governed by dvz /dt = (2 + Zi )0 (v) vz and dv/dt = 0 (v) v, respectively.
Combine these equations and show that for vT e < v < v0 their solution can be
written as vz = vz0 [v(t)/v0 ]2+Zi . Then, show that the current induced in the
plasma by one suprathermal electron over the time it takes for it to slow down
to the thermal energy of the background electrons is
%
qe vz0
1
Jz = qe dt vz " e/e
.
0 (v0 ) 5 + Zi
However, show that the sum over an isotropic distribution of such suprathermal
electrons yields no net current in the plasma. Next, consider the effect of a
small momentum input via radiofrequency waves at v0 = v0
ez that increases
the electron velocity to (v0 + v)
ez where v << v0 . Calculate the ratio of the
perturbed current J to the power (energy) input Pd from the wave needed to
produce this change and show it is given by
J
qe
4
=
v02 .
Pd
5 + Zi me 0e/e(v0 ) v0
Also consider wave momentum input in directions perpendicular to the initial
suprathermal velocity direction z with vz0 ,= 0; show that it too can induce current in the z direction (with reduced efficiency) and explain physically how this is

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

53

CHAPTER 2. COULOMB COLLISIONS

possible. Finally, show that when the current and power dissipated are normale/e
ized to ne qe vT e and ne me vT2 e 0 (vT e ), respectively, the normalized steady-state
current-drive efficiency for z-directed momentum input is given by
#
$2
J
4
v0
=
.
Pd
5 + Z i vT e
In what range of speeds is this type of current drive most efficient? How does
it compare to the normalized efficiency J/Pd for the usual ohmic current-drive
by an electric field with |E| << ED ? ///+

2.34 Show that for test electrons with energies much larger than the electron temperature in an electron-ion plasma with Ti Te that e " [2/(2 + Zi )]Se and that
the velocity friction and diffusion coefficients can be written to lowest order as
%v&e
t
%vv&e
t

e/e

(2 + Zi ) 0 v,

0 [(1 + Zi )(v 2 I vv) + (vT2 e /v 2 )vv].

e/e

How large are the most significant terms that have been neglected for the beam
electrons in Problem 2.12 /+
2.35 A hydrogen ice pellet is injected into a hot Maxwellian electron-proton plasma
with Te = Ti = 2 keV and ne = 5 1019 m3 . Assume the pellet doubles
the plasma density. Neglecting the energy expended in the ionization processes
( 30 100 eV), what is the temperature the hot plasma and the cold, pellet
produced plasma (at say 10 eV) will equilibrate to? Estimate the time scales
on which the electron and ion plasma components become Maxwellians at their
new temperatures, and equilibrate to a common temperature. //+
2.36 It is usually difficult to measure directly the ion temperature of hydrogenic ions
(protons, deuterons, tritons) in a hot plasma. However, it is often possible to
determine the temperature of trace amounts of impurity ions in hot plasmas
by measuring the Doppler broadening of the line radiation produced by the deexcitation of excited, highly ionized states of the impurity ions. Show that the
impurity temperature is close to the ion temperature in Maxwellian plasmas
with comparable electron (e), ion (i), and impurity (Z) temperatures as follows.
First, assume for simplicity that the impurities are heated only through Coulomb
collisions with the hot plasma electrons and the dominant, hydrogenic ions.
Then, write down an energy balance equation for the impurity species. Next,
show that in equilibrium since the impurity mass is much closer to that of the
hydrogenic ions than to that of the electrons, the impurity temperature can be
written as
4
5
Z/e
Z/i
TZ " Ti $ /
$
(Ti Te )
#
$#
$1/2 # $3/2
ne
me
Ti
= Ti
(Ti Te ).
ni Zi2
mi
Te
Finally, estimate the difference between the impurity and ion temperatures for
Ti = 10 keV, Te = 5 keV in a predominantly electron-deuteron plasma. //+
2.37 Using the formulas in Section 2.8, show that the impurity factor fie given in
(2.50) is correct. /+

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

54

2.38 Using the formulas in Section 2.8, show that the total electron collision frequency
e for a muliple ion species, impure plasma is as given in (2.44). /+
2.39 Using the formulas in Section 2.8, show that the impurity factor fi given in
(2.48) is correct. /+
2.40 Show that for a fast ion of mass mf and charge Zf e with vT i << vf << vT e
slowing down in a plasma composed of a mixture of ions of mass mi and charge
Zi e the slowing down is governed by (2.118), with the fast ion slowing down
time S unchanged, but that the critical speed vc is now given by

2 ni Zi2 /ne
3 3 2 ni Zi2 /ne
3 me
vc3
vT e

[Z]m vT3 e , [Z]m


.
4
(mi /me )
4 mf
(mi /mf )
i
i
Here, [Z]m is a mass-weighted effective Zi for energy transfer processes for an
impure plasma. For the parameters of Problem 2.29, what fraction of the alpha
particle energy is transferred to the ions and what fraction to electrons? //+
2.41 Estimate the distance traveled by a fusion-produced 3.52 MeV particle in slowing down in an infinite, homogeneous thermonuclear plasma as follows. First,
calculate the alpha particle energy loss rate per unit distance traveled (d/dz)
in terms of quantities derived in Section 2.9. Then, integrate to obtain the total
distance z the alpha particle travels in slowing down from its initial velocity to
the thermal velocity of the background plasma. Finally, estimate the distance
traveled for the parameters of Problem 2.29 using the formulas developed in
Problem 2.40. //+
2.42 Calculate the current driven in a tokamak plasma by the fast ions introduced by
energetic netral beam injection [T. Ohkawa, Nuclear Fusion 10, 185 (1970)], as
follows. First, consider introducing a beam of fast ions of density nf , and charge
Zf e with velocity Vf such that vc << Vf << vT e . Calculate the relative flow
Ve Vi induced by the beam ions for a plasma having ions of charge Zi and
density ni . Assume nf << ne and that the beam ions transfer their momentum
only to plasma electrons for simplicity. Show that the net current in the plasma
due to the three plasma components is given by J = nf Zf eVf (1 Zf /Zi ).
Explain physically why there is no current when the fast beam ions have the same
charge as the background ions, which is sometimes called a plasma shielding
effect. [Hint: The beam momentum input to the electrons is the same as the loss
of fast ion momentum by collisions with the Maxwellian electron background,
f /e
namely mf nf S Vf .] //+
2.43 For the case where there are no direct particle losses during fast ion slowing
down (cx ), show that the fraction of fast ion energy transferred to the
background plasma ions can be written as
+
,
(1 + x)2
2x 1
1
1
2
1
1

Gi = 2 ln
arctan
+
arctan
+
x
6 1 x + x2
3
3
3
3
in which x v0 /vc . //+

2.44 For the parameters of the Problem 2.29, calculate the fast ion slowing down
characteristics for a D-T fusion-produced alpha particle: a) the critical energy
c , and b) the total lifetime from birth to thermalization in the background
D-T plasma. Also, estimate the fraction of the alpha particle energy that will
be transferred to the background plasma electrons and to the plasma ions. /+

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 2. COULOMB COLLISIONS

55

2.45 In the early, 1970s experiments that injected energetic neutral beams into tokamak plasmas there was concern that charge-exchange of the injected fast ions
with neutrals in the plasma would cause the fast ions to be lost from the plasma
before they could deposit their energy in the background electrons and ions.
Consider 40 keV deuterium beam injection into a ne = 2 1013 cm3, Te = 1.3
keV, Ti = 0.5 keV electron-deuteron plasma. What are the critical energy c ,
slowing down time S and fast ion lifetime f for fast ions in this plasma?
In the absence of charge-exchange losses, what fraction of the fast ion energy is transferred to the backgound plasma electrons, and ions? Assuming
a charge-exchange cross section of cx = 7 1016 cm2 and a neutral density
of nn = 2 108 cm3 , what is cx at the initial fast ion energy? (Since cx v
is approximately constant below 20 keV per nucleon, cx is nearly independent
of energy.) How much does charge-exchange reduce the fractions of fast ion energy transferrred to the background plasma electrons and ions? Which transfer
fraction is affected the most? Why? /+
2.46 Energetic neutral atoms from neutral beams are absorbed in plasmas via the
atomic collision processes of electron ionization, proton ionization and charge
exchange. The ionization processes are ion /(cx + ion ) probable ( 30%
for the parameters of the preceding problem). They produce an electron whose
initial speed is approximately the same as the injected fast neutral atom but
whose kinetic energy is much lower. For the parameters of the preceding problem, what is the energy of such electrons? Since such electrons are born with
low energies and take energy from the backgound plasma as they are heated to
the plasma electron temperature, they represent an initial heat sink. Approximately how long does it take for these electrons to be heated by the plasma
to the background electron temperature of 1.3 keV? (Hint: See Problem 2.32.)
How does this time compare to the fast ion slowing down time S ? About how
long does it take for the injected energetic neutral beam to add net energy to
the plasma? //+
2.47 In the wet wood burner approach to controlled thermonuclear fusion [J.M.
Dawson, H.P. Furth and F.H. Tenney, Phys. Rev. Letters 26, 1156 (1971)] it is
proposed to obtain energy multiplication through fusion reactions of energetic
deuterons as they slow down in a background triton plasma. Show that the
energy multiplication factor F , which is defined as the ratio of fusion energy
produced to the initial deuteron energy 0 , can be written as
# $
# $
%
f (n ) v0 v3 dv (v) f (n ) v
F =
S
S
f
f
o
v 3 + vc3
0
0

in which f is the energy produced per fusion and f (v) is the speed- (energy-)
dependent fusion cross section. For D-T fusion with f = 22.4 MeV (17.6 MeV
from the reaction products and 4.8 MeV from assuming energy multiplication
by neutrons absorbed in a surrounding lithium blanket) and f v " 2.8 1022
m3 /s for 120 keV deuterons, find the minimum electron temperature at which
energy multiplication is possible. For this critical temperature, what is the
probability that a deuteron will undergo a fusion reaction during its slowing
down? //+

2.48 Consider the angular or perpendicular diffusion of a fast proton with energy 40
keV injected into a Maxwellian electron-proton plasma that has Te = Ti = 1
keV and ne = 3 1019 m3 . Estimate the time at which Coulomb collisions

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

56

CHAPTER 2. COULOMB COLLISIONS

scatter the velocity space angle of the fast proton through = v /v = 0.1
( 6 ). Compare this time to the fast ion slowing down time S defined in
(2.115), and discuss the physical reason why one of these times is much shorter
than the other. /+
2.49 Consider the proposition that an electric field E is applied to keep fast ions
with vT i << v << vT e from slowing down in a plasma. First, calculate the
momentum loss rate of the fast ion and find its minimum as a function of the
fast ion speed v. Next, calculate the minimum electric field [in terms of the
Dreicer field defined in (2.28)] required to prevent fast ions from slowing down.
Discuss the degree to which such an electric field would cause runaway electrons.
Finally, estimate the rates of perpendicular and parallel diffusion at the fast ion
speed at which the minimum momentum loss rate occurs, and discuss the effects
these processes might have on the proposed scheme (cf., Fig. 2.7). //+
2.50 Show that for a fast ion slowing down in the plasma described in Problem 2.40
the velocity friction and diffusion coefficients can be written to lowest order as
#
$
%v&f
vc3 v
= [Z]m + (Zeff + [Z]m ) 3
t
v
S
#
$
f
3
%vv&
2[Te + Teff (vc3 /v 3 )] vv
v
1
=
Zeff c3 (v 2 I vv) +
t
S
v
mf
v2
in which the angular scattering Zeff is defined in (2.43), the energy transfer or
mass-weighted [Z]m is defined in Problem 2.40 and the effective ion temperature
is defined by
Teff

1 2 ni Zi2 /ne
Ti .
[Z]m i (mi /mf )

How large is the most significant term that has been neglected in these approximate results and where does it contribute for fast ions slowing down for the
situation described in Problem 2.48? //+
2.51 Write a Monte Carlo type computer code for exploring the Coulomb scattering
of energetic test electrons in an electron-proton plasma. Use it to determine
numerically the answers to parts a) and b) of Problem 2.12. //+
2.52 Write a Monte Carlo type computer code for exploring the Coulomb collision
processes for fast ions slowing down in an electron-ion plasma. Use it to determine numerically the answers to the questions in Problem 2.48. ///+

DRAFT 14:53
August 25, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

Chapter 3

Structure of Magnetic
Fields
Many of the most interesting plasmas are permeated by or imbedded in magnetic
fields.1 As shown in Fig. 3.1, the magnetic field structures in which plasmas are
immersed are very diverse; they can also be quite complicated. Many properties
of magnetic fields in plasmas can be discussed without specifying a model for
the plasma. This chapter discusses the plasma-independent, general properties
(kinematics) of magnetic fields, the models commonly used to describe them
in plasma physics, and the coordinate systems based on them.
As indicated in Fig. 3.1, the generic structure of the magnetic field can be
open (ac and f) or closed (d,e). In open configurations the ends of the magnetic
field lines2 may intersect material boundaries (e.g., the earth in b), or be left
unspecified (e.g., in a, on the field lines in b that do not intersect the earth,
and in f). The magnetic field structure in closed configurations (d,e) is toroidal
in character or topology. That is, its magnetic field lines are topologically
equivalent (at least approximately) to lines on the surface of a torus or donut.
In most magnetized plasma situations the magnetic field has a nonzero value
and a locally specified direction throughout the plasma. Also, the flow of magnetic field lines penetrating a closed surface in the plasma often3 forms a bundle
1 In plasma physics when we say magnetic field we usually mean magnetic induction
field B both because for many plasmas embedded in magnetic fields the plasma-induced
currents are small and hence the magnetic permeability is approximately that of free space
(i.e., ! 0 ), and because most plasma calculations, which use the microscopic Maxwells
equations, assume that the charged particles in the plasma produce currents in free space
rather than doing so in a dielectric medium.
2 While magnetic field lines or lines of force do not in fact exist (at least in the sense
that they can be directly measured), they are very useful theoretical constructs for visualizing
magnetic fields.
3 However, closed magnetic flux surfaces do not exist in regions where the field lines are
chaotic. Also, there are sometimes null points of the magnetic field within the plasma for
example in the neutral sheet in the earths magnetosphere shown in Fig. 3.1b and along the
axis in the wiggler field for the free electron laser shown in Fig. 3.1f. In addition, certain

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

solar

wind

N
S

neutral!sheath
plasmasphere

a)!magnetic!mirror

b)!earths!magnetosphere

c)!screw!pinch

d)!tokamak

e)!stellarator

f)!free!electron!laser!wiggler!field

Figure 3.1: Examples of magnetic field configurations in which plasmas are


imbedded.

of magnetic field lines bounded by a magnetic flux surface within the magnetic
components of the magnetic field may have null points as well for example the projection
of the magnetic field in a screw pinch tokamak or stellarator in Figs. 3.1ce along the helical
pitch of a given magnetic field line.

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

configuration. When nested magnetic flux surfaces exist, they usually provide
the most natural magnetic-field-based coordinate system because most plasma
processes (charged particle motion, flows, transport) are much more rapid along
magnetic field lines and within flux surfaces than across them.
The key magnetic field variations are evident in the magnetic field configurations illustrated in Fig. 3.1. Namely, while magnetic field lines point in a given
direction at any specified point, they can curve and twist, and their density can
vary in space. The general properties of curves along a vector field such as the
magnetic field are summarized in Section D.6.
The first section of this chapter introduces simple models (quadratic and
sinusoidal magnetic well, and sheared slab magnetic field) that include the four
most important local magnetic field properties for plasma physics (namely, parallel and perpendicular gradients, curvature and shear). These simple models
will be used to explore the most fundamental effects of magnetic fields in many
areas of plasma physics throughout the remainder of this book. The second
section introduces the global magnetic field representations and magnetic-fieldbased coordinate systems that are used in modeling plasma processes on (longer)
time scales where charged particles travel significant distances along magnetic
fields. While the magnetic fields in which plasmas are imbedded are seldom
straight in Cartesian coordinates, one can develop coordinate systems in which
the magnetic field lines are straight. Such coordinates greatly facilitate analyses
of magnetized plasmas. The third section develops the basic ideas of magnetic
island structures that can form in the sheared slab model when a resonant
perturbation is added. The next three sections discuss the simplest forms and
properties of magnetic field coordinate systems for open (Section 3.4) and closed
(Sections 3.5, 3.6) magnetic field systems. Finally, Section 3.7 gives the general
forms of all the local differential propoerties of the vector magnetic field gradients, divergence, curvature, shear and torsion, and the general expansion of
the magnetic field B in terms of them.

3.1

Local Properties

The SI (mks) units for the magnetic field strength are webers/m2 ; thus, we can
think of the magnetic field strength as representing the number of magnetic
field lines (webers) per unit area (m2 ). Typically, the magnetic field strength
varies as we move along a magnetic field line. We can distinguish the effects
of variations in the magnetic field strength from the effects of changes in its
direction by representing the magnetic field as

B = B(x) b,

B/B,
with b

(3.1)

is the local unit


in which B |B| B B is the magnetic field strength and b
vector along B, both at the point x. Since there are no magnetic monopoles in
nature, a magnetic field must be divergence free. (Such a vector field is called
a solenoidal field.) Thus, using the representation of B given in (3.1) and the
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

vector identity (??), we must have


B(!) + B b.

0 = B = b

(3.2)

An equation governing the variation of the magnetic field strength B along any
field line can be determined by rearranging this equation to yield
B
B = B b,

b
!

(3.3)

in which ! is the distance along a magnetic field line. Hence, if the magnetic
field strength (number of field lines per unit area) is increasing (B/! > 0)
as one moves along a magnetic field line, the local unit vectors along magnetic
< 0); conversely, for a decreasing magnetic
field lines must be converging ( b
> 0).
field strength (B/! < 0) the field line unit vectors diverge ( b
We will often be interested in describing mathematically the parallel (%)
variation of the magnetic field strength B. Near a minimum in the magnetic
field strength along a magnetic field line the field strength B can be represented
by a quadratic approximation:
!
"
!2
Bqw = Bmin 1 + 2 , quadratic well (qw) model,
(3.4)
L!
in which at B = Bmin where ! = 0 we have B/!|!=0 = 0 and 2 B/ !2 |!=0 > 0,
and by definition
#
$
2 B $$
.
(3.5)
L!
2 B/!2 $B=Bmin

The characteristic scale length L! is the parallel distance over which the magnetic field strength doubles in this lowest order approximation.
The magnetic field strength often varies sinusoidally along a magnetic field
line. A convenient model for this variation is
%
&'
%
&(
Bmax Bmin
2!
Bsin (!) = Bmin +
1 cos
2
L!
= Bmin + B sin2

!
,
L!

sinusoidal (sin) model.

(3.6)

Here, Bmax is the maximum field strength along a field line which occurs at
! = L! /2 in this model, and B Bmax Bmin is the amplitude of the
variation of B along a field line within the periodicity length L! . The ! variation
of Bsin near its minimum can be represented by the parabolic well model in
(3.4) with L! = (Bmin /B)1/2 L! /. The ratio of the maximum to minimum
magnetic field strength along a field line is:
Rm
DRAFT 22:52
September 22, 2003

Bmax
B
= 1+
,
Bmin
Bmin

c
!J.D
Callen,

magnetic mirror ratio.

(3.7)

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

Mirror ratios range from values of order 2 to 10 or more for typical magnetic
mirrors (Fig. 3.1a) and the earths magnetosphere (Fig. 3.1b), to only slightly
greater than unity in toroidal devices (Fig. 3.1d,e) where the magnetic field
strength varies only slightly as we move along the helical magnetic field lines
from the outside to the inside of the torus. Note from (3.3) that at an extremum
(minimum or maximum) of the magnetic field strength where B/! = 0 the
local unit vectors along field lines are divergence-free they neither converge
nor diverge.
The magnetic field can also vary both in magnitude and direction
in directions perpendicular (transverse) to the magnetic field direction. The
sheared slab model, which we now discuss, approximates the local perpendicular
variations of typical magnetic field structures that are most important in plasma
physics. In it a local Cartesian coordinate system is constructed at a given point.
The z axis is taken to be along the magnetic field at the point where a magnetic
field line passes through the origin of the coordinate system. The x axis is
taken to be in the radial (across flux surface) direction in which the most
significant variations (in plasma parameters and in the density of magnetic field
lines) occur in the plane perpendicular to the magnetic field. The y axis is
taken to be in the azimuthal (or within flux surface) direction of least variation;
i.e., it is the ignorable coordinate, at least approximately. For example, for
a cylindrical magnetized plasma we anticipate mainly a radial variation in the
plasma parameters: for this case the sheared slab model x, y, z coordinates
would correspond to r r0 , r0 and z where r = r0 is the cylindrical radius of
the magnetic field line that passes through the origin of the sheared slab model.
The word slab in the title of the model indicates that only a thin radial (x)
slice of the magnetic configuration is being considered.
A local expansion of the magnetic field that captures its most important
perpendicular variations is
Bss

'%
&
(
z
x
x
= B0 1 +

ex +

ey ,

ez +
LB
RC
LS

sheared slab (ss) model,


(3.8)

in which B0 is the strength of the magnetic field (or density of magnetic field
lines) at the origin where x (x, y, z) = (0, 0, 0). Here, as indicated in Fig. 3.2,
the
ez term represents the lowest order magnetic field (the unity) and the perex term represents the
pendicular spatial gradient of its magnitude (1/LB ), the
ey term represents the differential
magnetic field curvature (1/RC ), and the
twisting (shear, 1/LS ) of the magnetic field lines. These fundamental magnetic
field properties will be explained and defined more precisely below and in the
following sections. [Torsion (uniform twisting see Section 3.7 below and D.6)
of magnetic field lines such as in a uniform helical twist of the field lines in the
screw pinch shown in Fig. 3.1c is not included in the sheared slab model because
the
ez vector is taken to be in a locally fixed rather than rotating direction.]
Since the model represents a Taylor series expansion of the magnetic field about
a given point, it is only valid for small distances from the origin |x/LB | << 1,
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

x0

x0

!"

d!

!"

B0

x(!)

Rc

x0

!"

d!

#"

y(!,!x)

Ls

Figure 3.2: Magnetic field line characteristics included in the sheared slab magnetic field model. Each sketch indicates the behavior of magnetic field lines
when only the indicated coefficient does not vanish.

|z/RC | << 1, |x/LS | << 1.

Calculating the magnitude of the magnetic field using B |B| = B B,


we find, to lowest order in the distance from the origin,
) 2
%
*&
%
&
x x2 z 2
x
x
+O
,
,
&
B
1
+
. (3.9)
Bss |Bss | = B0 1 +
0
2
LB
L2B L2S RC
LB

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

Thus, we identify
LB

B
=
dB/dx

d ln B
dx

&1
,

perpendicular B scale length,

(3.10)

in which the differential is to be evaluated at the origin of our local Cartesian


coordinate system. The gradient scale length LB is the radial (x) distance
over which the magnitude of the magnetic field would double in this linear
model. Hence, the 1/LB term in (3.8) represents the gradient in the magnetic
field strength (density of magnetic field lines) in the x direction (cf., Fig. 3.2c).
Henceforth, we will call this the perpendicular B or gradient B term.
The curvature of a magnetic field line can be determined as follows. First, we
propose that a coordinate function x(!) represents the x variation of a magnetic
field line as we move a distance ! along it. Then, the x component of the
curvature of the magnetic field line is defined as the second derivative of x(!)
along the field line:
curvature

d2 x(!)
.
d!2

(3.11)

For a magnetic field line near the origin of the sheared slab model coordinate
system, by geometry we have
Bx (!)
Bx (z)
z
dx(!)
&
&
=
d!
B
B0
RC
and hence
curvature

'
(
d Bx (z)
1
d2 x
&
=
d!2
dz
B0
RC

(3.12)

(3.13)

ex B is the x component of the vector magnetic field. The radius


in which Bx
of curvature RC of the magnetic field in the sheared slab model is the radius
of the circle that is tangent to and has the same curvature as the magnetic
field line that passes through the origin. Integrating (3.12) a short distance
(|z/RC | << 1) along the field line that passes through x = (x0 , 0, 0) yields an
equation for the field lines trajectory (cf., Fig. 3.2a) in the xz plane (to lowest
order d! & dz and ! & z near the origin):
x = x0 + z 2 /2RC ,

for y = constant,

(3.14)

which again shows that 1/RC measures the curvature of the field line.
The formal definition of the curvature vector for a vector magnetic field
is [see (??) in Section D.6]
B Bb

d2 x
)b
= RC ,
= (b
2
2
d!
RC

DRAFT 22:52
September 22, 2003

B field curvature vector.

c
!J.D
Callen,

(3.15)

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

Evaluating this expression for the sheared slab model magnetic field in (3.8),
we obtain (near the origin where d! & dz and |x| << LB )
=

'
(

(z/RC )
b
ex + (x/LS )
ey
1
&

ez +

ex x
ex .
&
!
z
1 + x/LB
RC

(3.16)

Thus, RC is the inverse of the (normal, x direction) curvature of the magnetic


field:
RC 1/|x |,

radius of curvature.

(3.17)

Note that the absolute value is needed because vectorially the radius of curvature
vector RC points in the opposite direction from the curvature vector: =
RC /Rc2 see Fig. ?? and Eq. (??) in Section D.6. Thus, for the sheared
ex , which
slab model the vectorial radius of curvature is RC /||2 = RC
points from the point x = (RC , 0, 0) to the origin.
The magnetic field line curvature vector can in general be written in a more
illustrative and useful form (for situations where currents flow in the plasma)
B/B and the magnetostatic Amperes law B = 0 J:
using b

=
=

)b
= b(

= b(B/B)

(b
b)

b[(1/B)B]
b(B)/B
= b(
ln B) + 0 JB/B 2
b
)]B + 0 JB/B 2 ,
(1/B)[ b(
(3.18)

in which the vector identities (??), (??), and (??) have been used in successive
steps. Defining
b),

(b
) = b(
b

gradient perpendicular to B, (3.19)

to represent the components of the gradient operator in directions perpendicular


to the magnetic field B, we can write the final form in (3.18) in general as
= ln B +

0 JB
,
B2

relation of curvature to B and J.

(3.20)

Near the origin of the sheared slab model, the


ex component of this equation
yields
x

0 Jy
1
1
1 dB
0 Jy
+
=
.
=
+
RC
B dx
B
LB
B

(3.21)

When there is no current in the


ey direction in the sheared slab model, we have
1/RC = 1/LB .
The shear in a magnetic field can be understood as follows. A magnetic field
line can rotate about the z axis because of torsion (twisting at a constant angular
rate) and shear (differential twisting) see Section D.6. As noted above, the
sheared slab model does not include torsion. The shear in the magnetic field

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

can be defined for our Cartesian coordinate system through the x derivative of
the y(!) coordinate variation along a magnetic field line:
'
(
d dy(!)
.
(3.22)
shear
dx
d!
For a magnetic field line near the origin of the sheared slab model coordinate
system, by geometry we have
By
x
dy(!)
&
&
.
d!
B0
LS
Thus, in the sheared slab model we have
% &
'
(
d By (x)
d dy
1
&
shear
.
&
dx d!
dx
B0
LS

(3.23)

The shear length LS is the linear extrapolation distance in the x direction over
which the magnetic field would differentially twist through an angle of one radian
(i.e., to where By = B0 ). Integrating (3.22) a short distance (|z/LS | << 1) along
the field line that passes through the point x = (x, y0 , 0) yields an equation for
its trajectory (cf., Fig. 3.2b) in the yz plane:
y = y0 + x z/LS ,

for x = constant,

(3.24)

which shows that 1/LS measures the differential twisting of the field lines out
of the the xz plane and hence the shear in the magnetic field lines.
is [see
The formal definition of the local shear in a vector field B B b
(3.151) below and (??) in Section D.6]

(b)
(b)
(B) (B)
=
,
2

B 2 ||2
|b|

local shear
(3.25)

in which is the gradient of an assumed magnetic flux function and for


the last form we have used (3.1) and vector identities (??), (??) and (??). For

ex =
b
our sheared slab model x, x =
ex and thus b
ez (x/LS )/(1 + x/LB ). Note that near the origin of the sheared slab model

ey
& 1 and |b|

geometry |b|
& 1. Thus, evaluating the shear definition in
(3.25) for the sheared slab model we obtain
ex ) & 1/LS .
&
ey (b

(3.26)

By construction, the magnetic field in the sheared slab model satisfies the
solenoidal or no magnetic monopole condition for a magnetic induction field,
i.e., Bss = 0. However, its curl (rotation) does not vanish:
'%
&
(
1
1
1

ez .
Bss = B0

ey +
(3.27)
RC
LB
LS
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

10

In equilibrium situations where the magnetostatic Amperes law B = 0 J


applies, the full generality of the sheared slab model is appropriate only if electrical currents flow in the plasma. For vacuum or very low plasma pressure
situations where no significant currents flow in a magnetized plasma, we must
have 1/RC = 1/LB and 1/LS = 0. The curvature (1/RC ) can deviate from the
inverse gradient length (1/LB ) only if electrical current flows in the y direction,
as indicated by both (3.21) and (3.27). Since for strong magnetic fields it is
harder for charged particles and hence plasma currents to flow across magnetic
fields compared to along them, the y component of the current is typically small
and usually 1/RC & 1/LB . Magnetic shear (1/LS ) is possible (in this torsionfree model) only if current flows in the z (magnetic field) direction. These points
will be made more quantitatively explicit in Sections 3.7, 5.3 and 20.1.
The parallel quadratic well, sinusoidal and sheared slab models represent the
most important spatial variations of the magnetic field around a given point.
Any given physical situation can be modeled with these models by specifying
the characteristic scale lengths for the local properties of the magnetic field:
parallel gradient B scale lengths L! and L! , perpendicular gradient B length LB ,
curvature radius RC and shear length LS . While these models provide suitable
lowest order local descriptions for most magnetized plasma situations, they
are not the most general magnetic field descriptions. In particular, they do not
allow for torsion or all the possible magnetic field variations in the y and z
directions. The most general local expansion of a magnetic field is discussed in
Section 3.7. Also, the local expansions do not in general provide global (i.e., valid
over all space) descriptions of the magnetic field. The remaining sections of this
chapter develop more complete, but correspondingly more complex, magnetic
field models.

3.2

Magnetic Field Representations and Coordinate Systems

In the preceding section we developed local Taylor series expansions of a magnetic field B about a given point. While these expansions are very useful for
understanding the local differential properties (gradients, curvature, shear) of a
magnetic field, in general they do not provide a global description of it. Charged
particles in plasmas move over long distances along magnetic field lines for most
time scales of interest. Also, they typically move much more rapidly along magnetic field lines than perpendicular to them; this causes the properties of a
magnetized plasma to be very anisotropic relative to the magnetic field direction. In order to develop compact descriptions of magnetized plasmas it is most
convenient to use coordinate systems based on the global structure of the magnetic field so-called magnetic field line or magnetic flux coordinate systems.
Magnetic flux coordinates are curvilinear coordinates that are chosen so that
the equation of a magnetic field line is a straight line in the chosen coordinates. They are the most useful coordinates because they facilitate separation

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

11

of plasma effects along and perpendicular to magnetic field lines. This section
discusses calculations of magnetic field lines, and magnetic field representations
and coordinate systems that describe the entire magnetic field structure.
The global structure of the magnetic field can in principle be obtained by
simply integrating the differential equations of a curve that follows a magnetic
field line. Defining x(!) to be the trajectory along a magnetic field line, the
vector dx(!)/d! that is locally tangent to the magnetic field is given by
$
dx(!) $$
d! $along

=
B

= b,
B

field line equation.

(3.28)

This is the fundamental definition of a magnetic field line that we will use
ey , and
ez projections
throughout the remainder of this book. Taking the
ex ,
of this fundamental field line definition, we obtain
dy
dz
d!
dx
=
=
=
.
Bx
By
Bz
B

(3.29)

Note that these field line differential equations can also be obtained from the
condition that a vector differential length d" along the magnetic field B must
be parallel to it: d"B = 0.
For simple magnetic field systems we can directly integrate the three independent equations in (3.29) to obtain a mathematical description of the magnetic field. For example, we performed such integrations for the sheared slab
model in the special cases of no shear and little perpendicular gradient B or
curvature see (3.14) and (3.24). For such systems the constants of integration provide labels for the magnetic field lines x0 and y0 for the two special
sheared slab model cases. However, it is often impractical or impossible to
obtain a global magnetic field description by directly integrating the equations
that describe a magnetic field line trajectory. For example, integrating the three
equations for the complete sheared slab model in (3.8) results in a set of three
interrelated, implicit equations for which a closed solution is not possible, except
in the vicinity of the origin (see Problem 3.7).
For a magnetic field in free space (i.e., in a vacuum), or in the limit where
the currents flowing in the plasma are negligible, the equilibrium Amperes law
becomes simply B = 0. This equation can be satisfied by writing the
magnetic field in terms of a scalar potential M :
B = M ,

vacuum magnetic field representation.

(3.30)

For this case the solenoidal (no magnetic monopoles) condition B = 0 becomes the Laplace equation
2 M = 0.

(3.31)

Methods for solving the Laplace equation in various geometries are available in
many books on electromagnetic theory and other areas of physics. For magnetized plasmas such solutions are useful mainly in vacuum regions outside the
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

12

plasma, or as the lowest order magnetic field structure for cases where currents
in the plasma do not significantly change the magnetic field. However, for many
important magnetized plasma situations the electrical currents flowing in the
plasma are significant, and in fact very important, in determining the structure
and even topology of the magnetic field. Thus, solutions of (3.31) for vacuum
magnetic fields are not always useful for magnetized plasmas and we must look
elsewhere for broadly applicable descriptions.
Like any vector field subject to the solenoidal condition ( B = 0), the
magnetic induction field B can be written in terms of a vector potential A:
B = A.

(3.32)

For example, an appropriate vector potential for the sheared slab model is
%
&
z2
x2
x2

ez ,
(3.33)

ey B0
Ass = B0 x +
2LB
2RC
2LS
as can be verified by substituting it into (3.32) and comparing the result to
(3.8).
Alternatively (see Section D.5), the magnetic field can be written as
B = ,

Clebsch representation,

(3.34)

in which (x) and (x) are scalar stream functions (i.e., functions that are
constant along the vector field B) since B = B = 0. Note that the
representations of B in (3.32) and (3.34) are equivalent if we define
A = ,

or A = ,

(3.35)

since using vector identities (??), (??), and (??), we have =


and = = . Note also that the vector potential A
and the stream functions , are somewhat arbitrary since they yield the same
magnetic induction field B under the gauge transformations A A + (x),
and + f1 () or + f2 () (but not both f1 , f2 simultaneously) in
which , f1 , and f2 are arbitrary scalar functions of the variables indicated.
While the stream functions , must be continuous, they can be multivalued
(e.g., they can involve angular or cyclic variables). For examples of and
stream functions, see Problem 3.7, which develops them for the sheared slab
model, and the following sections.
The Clebsch representation of the magnetic field can be used as a basis for
a coordinate system that represents the global magnetic field structure the
Clebsch magnetic coordinate system. Along magnetic field lines, which follow
the curve given by (3.28), we have d/d! = (dx/d!) = (B )/B = 0
and similarly d/d! = 0. Thus, magnetic field lines lie within (x) = constant
and (x) = constant surfaces. Further, since points in the direction of
(and is equal to) B, the intersection of the , surfaces defines a given magnetic
field line. Hence, and are labels for a particular magnetic field line.
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

13

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

Because the , stream functions label magnetic field lines in the plane
perpendicular to B, they can provide curvilinear coordinates perpendicular to
the magnetic field. There is no obvious choice for the coordinate along the
magnetic field. From physical considerations it is convenient to choose the
length ! (measured from some suitable surface) along field lines. [However,
other coordinates along the magnetic field are often used, e.g., dM = Bd!
for the vacuum magnetic field in (3.30)]. Unfortunately, the , , ! coordinates
are in general not orthogonal and not available in closed form solutions. These
complications plus their possible multivaluedness make them an awkward choice
as the basis for a magnetic-field-based coordinate system. However, because
of their simplicity and generality they are often useful for proofs concerning
equilibrium, stability and transport properties of magnetized plasmas.
Magnetic flux surfaces usually provide a better basis for developing magneticfield-based coordinate systems for plasma physics. The magnetic flux through
a surface S encompassed by a closed curve C is in general defined by
=

++

dS B =

++

dS A =

d" A,

magnetic flux,

(3.36)

in which we have used Stokes theorem (??) in the last step. In this book we
will use a capital letter to indicate the total magnetic flux in its normal units
(webers), and a small Greek letter to indicate a magnetic flux component that
has been normalized in some way (e.g., often = /2). Since magnetic flux
surfaces encompass the bundle of magnetic field lines within the surface S, they
must satisfy
B = 0,

magnetic flux surface condition.

(3.37)

Thus, , which by definition [see (??)] is normal to the flux surface (x), is
orthogonal to the magnetic field B and hence to its field lines. That is, magnetic
field lines lie within (x) = constant surfaces.
For a Clebsch coordinate system with A = and a closed contour C ,
the magnetic flux becomes
++
,
,
dS() B =
d"() =
d .
(3.38)
=
S

Here, the subscript is placed on and a argument is given for dS to indicate


that this magnetic flux will represent (see below) a magnetic field component
orthogonal to both the and coordinates. (For example, dS .)
Because the Clebsch representation is general, we will use this form of the magnetic flux both as a description of the complete magnetic field, and for individual
magnetic field components. To obtain the functional dependence of a magnetic
flux function it is often simplest to calulate it on a surface where it can be
evaluated easily and then extend it to other spatial positions by mapping the
magnetic field lines it encompasses to the new positions.
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

14

Magnetic flux surfaces can be constructed easily for magnetic configurations


or magnetic field components in which there is symmetry (i.e., no dependence
on a coordinate) in a direction perpendicular to the magnetic field. Then,
we choose to be that symmetry coordinate, and the magnetic flux and the
corresponding vector potential become
,

, for symmetry in .
(3.39)
= d, A = = d

When there is symmetry in the direction, the magnetic field component produced by the component of the vector potential in the direction can be
represented in terms of the corresponding magnetic flux by
%
&

, for symmetry in .
(3.40)
B = A = d

This component of the magnetic field is labeled with a vector cross product
subscript () because it is orthogonal to both the symmetry coordinate and the
flux coordinate directions: B = 0, and B = 0. Note that is
clearly a magnetic flux function since it satisfies (3.37). As a simple example of
how to directly use these formulas for a single component magnetic field, Section
3.4 develops the magnetic flux (and Clebsch) coordinates for an axisymmetric
magnetic mirror.
These formulas can be used to develop magnetic flux coordinates for the
sheared slab model as follows. In the absence of magnetic shear (i.e., for 1/LS
0), the sheared slab model is symmetric in the y direction. For this case, the
dominant or main magnetic field
- component in the sheared slab model can
be calculated by taking = y, d = y0 . Then, we use the rectangular
surface in the z = 0 plane specified by (see Fig. 3.3a) 0 x x0 and 0
y. y0.for calculating the magnetic flux in the z direction to yield z |z=0 =
x0
y
dx 0 0 dy Bz = (x0 + x20 /2LB )y0 B0 at z = 0. This magnetic flux is extended
0
to other (small) z values using the field line label x0 = x z 2 /2RC from (3.14)
to yield:
%
&
x2
z2
y

y,
y0 B0 & x y0 B0 , Ay
z & x +
2LB
2RC
y0
% &
'%
&
(
z
x
z
Bmain = Ay =

ey & B0 1 +

ez +

ex . (3.41)
y0
LB
RC
To determine a similar magnetic flux form for the auxiliary magnetic shear
component in the sheared slab model, we consider the case where the perpendicular gradient in B and curvature are absent (i.e., 1/LB = 0 and 1/RC = 0),
and the field line label simplifies -to x0 = x. Then, there is symmetry in the
z direction, and we take = z, d = z0 . Using the rectangular surface in
the y = 0 plane specified
(see Fig. 3.3b) 0 x x0 and 0 z z0 , we
.x
.by
z
obtain y |y=0 = 0 0 dx 0 0 dz By = (x20 /2LS )z0 B0 . (The y magnetic flux
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

15

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS


y
y

y0
x0

x
x0

By(x)

z0

z
a)!!main!magnetic!flux!ym

b)!!auxiliary!magnetic!flux!y*

Figure 3.3: Geometry of the surfaces through which the (a) main (z ) and (b)
auxiliary (y ) magnetic fluxes are calculated for the sheared slab magnetic field
model.

ex dx
ez dz =
ey dxdz.) Using x0 = x,
is negative because dS d"x d"z =
this yields
y
Az
z,
z0
% &
y
x
= Az =

ez = B0

ey .
z0
LS

y =
Baux

x2
z0 B0 ,
2LS

(3.42)

The total magnetic field in the sheared slab model can be represented in
terms of its y and z magnetic flux components by adding these two results:
Bss = Bmain + Baux = (z /y0 )y + (y /z0 )z.

(3.43)

Neglecting terms of order x2 and z 2 , the two components in (3.43) can be


combined into a single form B & B0 (x+x2 /2LB z 2 /2RC )(y x z/LS ) =
z y0 using (3.24), which is in the Clebsch form given in (3.34). However,
in general the two magnetic flux forms cannot be combined into a single Clebsch
form. For the sheared slab model the natural curvilinear coordinates near the
origin that can be deduced from this magnetic flux model of the magnetic field
are z , y0 y x z/LS and ! & z. Note that despite the presence of magnetic
shear, curvature and a perpendicular gradient of B, magnetic field lines are, as
desired, straight to first order in this magnetic flux coordinate system: dz /d! =
0, dy0 /d! = 0, and d! & dz along field lines.
Many physically relevant situations are more complicated, either because
they are fully three-dimensional and have no symmetry direction (e.g., the outer
parts of the earths magnetosphere in Fig. 3.1b and the stellarator in Fig. 3.1e),
or because there is a magnetic field component in the symmetry direction(s)
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

16

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

(e.g., the screw pinch in Fig. 3.1e and the tokamak in Fig. 3.1d). When there
is more than one magnetic field component and one of the components is in a
symmetry direction, the magnetic induction field B can be written in terms of
the magnetic flux components associated with a main (parallel to a symmetry
or periodicity direction) and an auxiliary (perpendicular to the dominant symmetry direction, or due to shear, torsion) component of the magnetic field. Each
magnetic field component can be written in terms of the relevant magnetic flux
in a Clebsch form using (3.40).
In general, representations of B fields can always be constructed with two
magnetic flux functions. They are quite useful in plasma physics. A single or
total Clebsch form can be developed from them whenever the two flux functions
are single valued functions of each other, which happens when they represent
configurations with closed, nested toroidal magnetic flux surfaces. Examples of
such systems include axisymmetric toroidal configurations (see Section 3.6) and
some regions of stellarators.
For toroidal magnetic configurations with helical magnetic field lines there
are two natural cyclic coordinates: the toroidal (long way around the torus) and
poloidal (short way) angles and . For the moment these will be arbitrarilydefined angles; they are only required to span their respective spaces. Then, in
analogy with (3.41) and (3.42), it can be shown in general that the magnetic
field can be written in the form of toroidal (tor) and poloidal (pol) magnetic
field and flux components:
B = Btor + Bpol = (tor /2) + (pol /2).

(3.44)

The natural sign of the poloidal magnetic flux pol would be negative for this
geometry because of the choice of , , as a right-handed set of coordinates;
however, by convention its sign is changed in this definition. The magnetic axis
(origin) for the poloidal angle coordinate is defined to be the line on which
Bpol (pol /2) vanishes.
In regions where a set of nested toroidal magnetic flux surfaces exist, the
poloidal flux function is a single-valued (monotonic) function of the toroidal
flux function and hence can be written in terms of it: pol = pol (tor ). Then,
the poloidal and toroidal angles can be modified ( f and f ), so that
magnetic field lines are straight in them (hence, the f subscript indicating
proper flux surface coordinates). (See Section 3.6 for the development of such
straight-field-line coordinates for axisymmetric toroidal configurations.) Thus,
for toroidal configurations with nested flux surfaces, the magnetic field in (3.44)
can be written compactly in the Clebsch form
B =

tor
2

&

DRAFT 22:52
September 22, 2003

/
0 1 2
f
f ,
2

c
!J.D
Callen,

toroidal flux surfaces B field,


(3.45)

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS


where we have defined
%
&
dpol
(tor ) 2
,
dtor

rotational transform angle (degrees),

17

(3.46)

which is the slope (df /df ) of the magnetic field lines in the f f plane. Here,
is the small Greek letter iota; it is divided by 2 in many formulas to represent
the angle of field line rotation (per toroidal transit) in radians. For this model
magnetic field we identify the Clebsch coordinates as = tor /2 and =
f (/2) f . Along magnetic field lines we have d = dtor /2 = 0 and d = 0
or df = (/2) df = f = (/2) f + constant. Thus, magnetic field lines are
straight in the tor =constant, f f plane. For such toroidal configurations the
natural magnetic field curvilinear coordinates are those based on the magnetic
flux coordinates tor , f , and f , which unfortunately are not usually orthogonal.
Nonetheless, since /2 is typically not a rational number (ratio of integers,
see Section 3.6), the magnetic flux coordinates usually provide a more useful
description than the Clebsch coordinates because of the multivaluedness of
the coordinate in f and f and because ! (or some other coordinate along field
lines) is not one of the natural coordinates of the magnetic field description.

3.3

Magnetic Islands

This section will explain how an error magnetic field can create a magnetic island
in a sheared magnetic field model it is yet to be written and inserted. The
main point of this section will be to show that when a resonant magnetic field
x sin ky is added to the sheared slab model
x = B
perturbation of the type B
x /kB0 )1/2 and to elucidate
it produces a magnetic island of width w = 4 (LS B
various properties of field lines in and around the magnetic island structure.

3.4

Open Magnetic Configurations*

There are many types of open magnetic configurations: a cylindrical column


of magnetized plasma, magnetic mirrors (Fig. 3.1a), the earths magnetosphere
(Fig. 3.1b), the interplanetary magnetic field, solar flares, cusps (produced by
pairs of mirror coils in which the coil currents flow in opposite directions),
and so-called divertor regions on open field lines that are outside the closed
flux surfaces in toroidal configurations. The simplest and conceptually most
important open configurations are of the axisymmetric magnetic mirror type,
as shown in Fig. 3.4.
We consider first an axisymmetric magnetic mirror composed of two identical
current-carrying solenoidal coils separated by a distance L, as shown in Fig. 3.4a.
This simple mirror is an important paradigm for discussing many effects of
geometry on magnetized plasmas. Since there is symmetry in the azimuthal ()

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

18

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

-!2

-!1

!0

!1

!2

2
1

!
B

0!=!0

!"

r0

|B|

|B|

Bmax

2Bmin
Bmin
-L!/2

-L"

L"

L!/2

-60

60

Figure 3.4: Two fundamental types of axisymmetric open magnetic field configurations: a) (on left) axisymmetric or simple magnetic mirror; and b) (on right)
a dipole magnetic field such as that due to the earths magnetic dipole.

direction and no component of B in this direction, we can construct a Clebschtype magnetic flux coordinate system using (3.39) and (3.40):
B =

m
2

&

axisymmetric mirror (m) magnetic field.

(3.47)

Here, we have taken = and used (3.39) to


- identify the magnetic flux for
an axisymmetric magnetic mirror as m = d = 2. A vector potential
e . This
that produces this magnetic field is A = (m /2) = (m /2R)
representation can also be used to describe the bumpy cylinder magnetic field
produced by a set of solenoidal coils confining a cylindrical column of magnetized
plasma (see Problems 3.1 and 3.12).
The magnetic field structure in an axisymmetric magnetic mirror is one of
the simplest nontrivial magnetic configurations. In particular, as can be seen
from Fig. 3.4a, because of the axisymmetry, it has no gradient of B or curvature
in the azimuthal () direction, Also, it has no shear or torsion. However, there
are axial (and parallel) and radial gradients of B in an axisymmetric mirror.
Further, the magnetic field lines have normal (see Section D.6) curvature (N
m -= 0). When the sheared slab model in (3.8) is used to describe the
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

19

local magnetic field in a simple mirror, we make the following associations:


(x, y, z) (m , , !), 1/RC = x m /|m |, 1/LB = (1/B)(dB/dx)
|m |(1/B)(B/m ), 1/LS 0.
In a simple mirror the magnetic field strength varies significantly along magnetic field lines. It is smallest at the midplane (Z = 0) between the mirrors and
maximum in the mirror throats (Z = L/2) of the two coils. Since the variation of |B| along magnetic field lines is approximately sinusoidal between the
mirror coils, it is commonly represented by the sinusoidal model in (3.6). Near
the midplane at ! = Z = 0, to lowest order the variation of the magnetic field
strength is quadratic in ! and can be represented by the quadratic well model
(3.4). The mirror ratio Rm Bmax /Bmin increases with minor radius R from
its minimum on the axis of symmetry (R = 0). When the mirror coils are
separated by approximately their diameters, its on-axis value is about 23.
The magnetic flux coordinate system m , , ! for a simple mirror can be
related to a cylindrical coordinate system R, , Z constructed about the symmetry axis of the magnetic mirror. For simplicity we define ! = 0, Z = 0 at the
midplane between the two mirror coils. We calculate the relation between the
distance ! along a field line and the axial cylindrical coordinate Z as follows.
First, we take the dot product of the field line equation (3.28) with
eZ Z
to obtain
BZ
dZ
=
.
d!
B

(3.48)

e B = 0) and
Since there is no azimuthal magnetic field component (B
near the axis of symmetry (R = 0) we can see from Fig. 3.4a that BR << BZ ,
we have
3
2 & B [1 + (1/2)(B 2 /B 2 ) + ].
(3.49)
B = BZ2 + BR
Z
R
Z
Now, the cylindrical coordinate form of B = 0 is

BZ
1
(RBR ) +
= 0.
R R
Z
Integrating this equation over a small distance R at constant Z away from R = 0
where BR = 0 (by axisymmetry) assuming that, as will be demonstrated below,
BZ depends only weakly on R, yields
BR &

R BZ
R B
R!
&
& 2 Bmin .
2 Z
2 !
L!

(3.50)

Here, we have anticipated from (3.49) that B & BZ and ! & Z near R =
0, Z = 0, and in the last form we have used the quadratic well approximation
of (3.4). The radial magnetic field component BR is nonzero and negative to
provide the needed (for B = 0) convergence (dR/d! BR < 0) of the field
lines as the magnetic field strength increases away from the mirror midplane

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

20

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

(B/! > 0) see discussion after (3.3). Using (3.50) in (3.48), the length !
along a magnetic field line is given for small R (<< L! ) and ! (<< L! ) by
2
1 BR
d!
= 1+
+
dZ
2 BZ2

! = Z[1 + R2 Z 2 /6L4! + ],

eR (RZ 3 /3L4! + ).
! =
eZ (1 + R2 Z 2 /2L4! + ) +

(3.51)

Note that for R -= 0 the distance ! along field lines is longer than the axial
distance Z, and that this lengthening effect increases with the cylindrical radius
R. Note also that for this simple mirror ! does not point in the same direction
as B since the coefficient of
eR in ! is positive while BR < 0.
The total magnetic flux m within a cylindrical radius R can be determined
approximately at the Z = 0 plane by neglecting the slight variation of B with
R, and then extended along field lines using R R(!) and B(Z = 0) B(!):
++
+ 2 + R
dS B =
d
R% dR% B & R2 B(Z = 0) = R2 (!)B(!).
m
Z=0

(3.52)

The gradient of m , which defines one of the directions in the magnetic flux
coordinate system, is
eR + (RZ/L2! )
eZ ].
(m /2) & BR R + (R2 /2)(B/Z)Z & BR [

/R in (3.47) yields the desired magnetic


Using this result together with = e
field direction and magnitude variation along field lines for an axisymmetric
mirror near R = 0, Z = 0.
The magnetic flux within a given bundle of magnetic field lines is conserved
(since B m = 0) as we move along the field lines and the magnetic field
strength varies. Thus, the radius R(!) of a given magnetic flux surface (or field
line) can be determined from (3.52),
#
#
m
Bmin
= R(0)
, radius of flux surface.
R(!) &
(3.53)
B(!)
B(!)
Hence, the radius of a flux surface varies inversely with the square root of the
field strength flux surfaces get smaller in radius R as we move toward the
mirror throats.
The normal (m or radial direction) curvature of the magnetic field lines
can be obtained from the second derivative of R(!) along a field line: R
d2 R(!)/d!2 . Near the axis of symmetry and midplane of a simple mirror it is
given by
R & R/L2! ,

for R, |Z| << L! .

(3.54)

[This result can also be obtained from the definition R d(BR /B)/d! from
(3.15) see Problem 3.11]. Thus, as is obvious physically from the axisymmetric magnetic mirror geometry, the radius of curvature RC 1/|R | is infinite on
the symmetry axis (R = 0), but is finite for R -= 0 and decreases as R increases.
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

21

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

The variation of |B| in the radial (R) direction can be estimated from the
magnetic field curvature as follows. First, we recall that for small R, Z the
magnetic field can be expanded as indicated in (3.49). Next, we assume that
the plasma electrical current in the direction is small and can be neglected.
Then, the component of the equilibrium (/t 0) Amperes law becomes
0 =
e B =

BR
BZ
ln BZ

& BZ (R
).
Z
R
R

Thus, as we could also have deduced from (3.18), we have


!
"
R
Z2
R2
ln BZ
& R & 2
=
BZ & Bmin 1 + 2
.
R
L!
L!
2L2!
Using this result in the expression for |B| in (3.49), we find
|B| & Bmin

4!

!2
1+ 2
L!

"!

R2
1
2L2!

"

+ .

(3.55)

(3.56)

Hence, as can be discerned by looking at the density of the field lines sketched
in Fig. 3.4a, there is a saddle point in the magnetic field strength at the center
(R = 0, Z = ! = 0) of the simple mirror |B| increases along field lines
( 2 B/!2 > 0 near |Z| = 0), but decreases radially (B/R < 0, for R -=
0). Within the axisymmetric model of the magnetic mirror field, |B| always
decreases with radius R; hence the region near R = 0, Z = 0 is a magnetic hill
radially, but a magnetic well axially. It will turn out (see Chapter 21) that for
macroscopic plasma stability we need to place the plasma in a global magnetic
well (B/R > 0, 2 B/Z 2 > 0). A minimum-B or magnetic well mirror
configuration can be created by adding nonaxisymmetric, multipolar magnetic
fields that are produced by currents in alternating directions in a set of axial
wires (Ioffe bars) outside the mirror coils (see Section 21.1).
Next, we consider the axisymmetric magnetic field generated by the earths
magnetic dipole, as indicated in Fig. 3.4b. Since the electrical currents in the
plasma near the earth are too weak to significantly affect the magnetic field,
we need only calculate the vacuum field induced by the earths dipole magnetic
ez . The magnetic potential d induced by a point magnetic
moment E Md
dipole is given by ({0 /4} 1 for mks cgs units)
6 7 x
0
E
, magnetic potential for dipole field.
(3.57)
d =
4
|x|3
Using the spherical coordinate system shown in Fig. 3.4b, outside the earth
(r > RE ) the magnetic potential becomes
d =
DRAFT 22:52
September 22, 2003

6 7 M
6 7 M sin
0
d ez x
0
d
=

4
|x|3
4
r2
c
!J.D
Callen,

(3.58)

Fundamentals of Plasma Physics

22

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

in which |x| = r is the distance from the center of the earth. Here, we have used

ez x = z = r cos = r sin in which = /2 is the angle characterizing


the latitude from the equatorial plane ( = /2, = 0).
Evaluating the components of B = d in r, , spherical coordinates,
we obtain
6 7 2 M sin
6 7 M cos
0
d
0
d
,
B
=
, B = 0, diplole field.
Br =

4
r3
4
r3
(3.59)
The B component vanishes because of the axisymmetry about the earths
magnetic axis. The total magnetic field strength is thus given by
B =

3
6 7 M (1 + 3 sin2 )1/2
0
d
Br2 + B2 =
,
4
r3

(3.60)

which shows that the magnetic field strength increases with latitude and
decreases with radial distance (as 1/r3 ).
The magnetic flux d for a dipole (subscript d) magnetic field can be calculated from the magnetic field penetrating downward through a disk in the z =
constant plane that extends radially outward from r to infinity using dS(z)
ez ) = B cos :

ez , and B (
d =

++

dS(z) B =

r% dr%

d cos B =

6 7 2M cos2
0
d
.
4
r
(3.61)

er
The direction of dS(z) and sign of d were chosen so that d is in the
(radially outward) direction at = 0. The variation of the radius of a field
line as changes can be obtained from the constancy of the magnetic flux d
along field lines: r() = r0 cos2 in which r0 is the radius of the field line in
the equatorial plane. Using this field line result in (3.60), we find that along a
magnetic field line |B| (1 + 3 sin2 )1/2 / cos6 , which increases rapidly away
from the equator ( = 0) see Fig. 3.4b. Near the equatorial plane the
magnetic field strength can
be modeled by the quadratic well model of (3.4)
with ! & r0 and L! = ( 2/3)r0 (see Problem 3.14). Since B/! > 0 for
< 0) as we move
! > 0, magnetic field unit vectors converge (Br < 0, b
along field lines vertically, above and away from the equatorial plane, toward
the earths polar regions where the magnetic field strength is largest.
For a Clebsch-type magnetic flux representation of the dipole magnetic field
we take and /2, and thus have
B=

d
2

&

dipole magnetic field.

(3.62)

That this form reproduces the field components in (3.59) can be shown using
er
e =
e ,
e
e =
er (because = /2 ). Note
=
e /(r cos ),
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

23

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

also that the dipole magnetic field can be represented by the vector potential
e .
A = (d /2) = (d /2r)
Like the simple mirror, the earths dipole magnetic field has no shear or
torsion. However, there is normal (radial) curvature and, since this is a nearly
vacuum field, a concomitant radial gradient of |B|. Using (3.60) and J & 0 in
(3.20), we find for all
r =

3
ln B = ,
r
r

LB = RC =

r
,
3

curvature of vacuum dipole field.


(3.63)

Note that for the dipole field the radius of curvature is independent of latitude
and equal to its obvious value of 3/r in the equatorial plane ( = 0). When
the sheared slab model in (3.8) is used to describe the local magnetic field in
the earths dipole field, we make the following associations: (x, y, z) (, , !),
1/LB r , x r , and 1/LS = 0.
Since mathematical descriptions of nonaxisymmetric open magnetic configurations usually depend on the specifics of the particular case, we will not
develop any in detail. While the characteristics of particular open magnetic
configurations can be quite important for specific effects, the lowest order or
most fundamental properties of open configurations are usually dominated by
the open rather than closed nature of the field lines, the magnetic mirrors along
B, and the B and curvature of the field lines. These latter properties are all
included in the axisymmetric models developed above. Thus, the axisymmetric
simple mirror or dipole field models provide appropriate lowest order magnetic
field models for all open configurations.

3.5

Screw Pinch Model*

There are a number of types of axisymmetric toroidal magnetic field configurations used for plasma confinement: tokamaks (Fig. 3.1d), spherical tokamaks,
spheromaks and reversed field pinches devices whose interrelationships are
discussed at the end of this and the next section. The paradigm for the axisymmetric toroidal class of configurations is the tokamak, both because it is
the simplest axisymmetric toroidal magnetic configuration with two magnetic
field components, and because so many experimental tokamaks have been built
and operated worldwide in the pursuit of the magnetic confinement approach
to controlled fusion. In turn, the tokamak magnetic geometry is often approximated by a periodic cylinder (see Fig. 3.1c), which is called the screw pinch
model and the focus of this section. In this section and the following one we
develop the screw pinch and axisymmetric toroidal models in general, and then
indicate the lowest order tokamak forms in the usual large aspect ratio (thin
donut) expansion after approximate equalities (&). The use and forms of these
general magnetic field structures for other axisymmetric toroidal configurations
are discussed at the end of the sections: reversed field pinches at the end of this
section, and spherical tokamaks and spheromaks at the end of the next section.
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

24

Figure 3.5: Screw pinch model of tokamak magnetic field geometry. The axial
periodicity length 2R0 represents the periodicity of the tokamak in the toroidal
direction.
The key parameter that describes the degree of toroidicity in all toroidal
magnetic configurations is the reciprocal of the aspect ratio. The aspect ratio A
is defined as the ratio of the major (R0 ) to minor (r) radius of a given magnetic
flux surface in the torus. The degree of toroidicity in toroidal configurations is
indicated by the parameter
0

r
1
= ,
R0
A

inverse aspect ratio.

(3.64)

This is a small number for magnetic flux surfaces inside most standard tokamaks whose aspect ratios at the plasma edge typically range from 2.5 to 5. Thus,
it will be used as an expansion parameter in the analysis of tokamak magnetic
field systems.
There are two classes of intrinsically toroidal effects in tokamaks that need
to be taken into account for small but finite 0. First, there are the effects due
to the toroidal curvature: the toroidal curvature of the magnetic field lines and
the differences in the magnetic field strength on the inner (small R) and outer
(large R) sides of the torus [see Eq. (3.110) in the next section]. Second, and
most importantly, there is the double periodicity of the system in the toroidal
(long way around the torus) and poloidal (short way) angle variables and .
In the screw pinch (periodic cylinder) model of the tokamak the double
periodicity is taken into account, but the toroidal curvature effects are neglected.
This model uses an r, , z cylindical geometry, as indicated in Fig. 3.5. In the
screw pinch model, r reresents the minor radius (or flux surface label) and
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

25

represents the poloidal angle in the tokamak. The tokamaks periodicity in the
toroidal angle is modeled by requiring periodicity in the axial coordinate z
over the toroidal length of the torus, 2R0 . Thus, the axial distance z in the
screw pinch is associated with the tokamak toroidal angle through
(3.65)

z = R0 .

The magnetic field in a tokamak has two components. The main, toroidal
(, z) magnetic field Btor is produced by electrical currents flowing poloidally
mainly in coils wrapped poloidally around the torus, but also within the plasma.
The smaller (for 0 << 1) poloidal () magnetic field Bpol is produced by the
toroidal component of current flowing in the plasma. In the screw pinch model
both components can depend on the minor radius r, although the variation
of |Btor | with r is weak for typical tokamaks. Thus, the magnetic field in a
tokamak is modeled by
B = Btor + Bpol Bz (r)
ez + B (r)
e ,

screw pinch model field.

(3.66)

Note that in the screw pinch model there is symmetry in the z, directions and
magnetic field lines lie on constant radius (r) surfaces (
er dx/d! = Br /B = 0).
The poloidal magnetic field is related to the axial component of the current
density J through the axial component of the equilibrium Amperes law:
ez J

ez B = 0

1 d
[ rB (r) ] = 0 Jz (r).
r dr

(3.67)

Integrating this equation using the boundary condition (by symmetry) that
B = 0 at r = 0 yields
+
0 r % %
0 Iz (r)
r dr Jz (r% ) =
, poloidal magnetic field,
(3.68)
B (r) =
r 0
2r
..
dS(z) J is the axial current flowing within a radius r.
in which Iz (r)
Similarly, the radial variation of the toroidal magnetic field Bz is related to
the poloidal current density through
e B = 0 J = dBz /dr = 0 J ,
which upon integration using the boundary condition that the currents in the
external poloidal coils and the plasma produce a toroidal magnetic field strength
of Bz (0) B0 on the axis (r = 0) yields
(
'
+
0 r %
%
dr J (r ) & B0 , toroidal magnetic field. (3.69)
Bz (r) = B0 1
B0 0
In order to determine the radial dependence of Bz , we need a specific plasma
model for the poloidal current density J . However, as indicated by the approximate equality, the magnetic field induced by the poloidal current in a tokamak
is usually small because the helical pitch [see (3.73) below] of the field lines
is small, and because the plasma-pressure-induced currents are small for low
pressure plasmas.
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

26

The magnetic fluxes associated with the toroidal and poloidal magnetic fields
in the screw pinch model can be determined by calculating the magnetic fluxes
in the z and symmetry directions (see Fig. 3.5):
++
+ r
dS(z) Btor = 2
r% dr% Bz (r% ), toroidal magnetic flux, (3.70)
tor
0
++
+ r
dS() Bpol = 2R0
dr% B (r% ), poloidal magnetic flux. (3.71)
pol
0

The screw pinch magnetic field (3.66) can be written in terms of these magnetic
fluxes using (3.40) or A = (tor /2) (pol /2)(z/R0 ):
B = (tor /2) + (z/R0 )(pol /2)
= (tor /2)[ (/2) (z/R0 )]

(3.72)

in which we have used the definition of = (r) in (3.46). The last form is a
Clebsch representation with = tor /2 and = (/2) (z/R0 ). For this
Clebsch representation, the equation for a magnetic field line is d = 0 =
tor (r) = constant = r = constant and d = 0 = d = (/2) dz/R0 =
= z (/2R0 ) + constant. Thus, the magnetic field lines in a screw pinch lie
on r = constant surfaces and are naturally straight in the z plane with a
constant helical pitch (see Fig. 3.1c):
(r)
d
=
,
dz
2R0

helical pitch of field lines.

(3.73)

Note that the screw pinch model magnetic field is in the toroidal flux form of
(3.45) with the straight field line coordinates identified as f and f z/R0 .
It is customary to characterize the inverse of the pitch of the helix of magnetic
field lines in a tokamak by a global measure (see Fig. 3.8 and discussion in next
section) which is the number of toroidal (or axial periodicity length) transits of
a magnetic field line per poloidal transit ( increasing from 0 to 2):
# toroidal transits of a field line
, toroidal winding number
# poloidal transits of a field line
. 2R0 /(/2)
dz/2R0
2
dtor
r Bz (r)
=
=
.
(3.74)
= 0
=
. 2
(r)
d
R
pol
0 B (r)
d/2
0

q(r)

The q value is also known as the safety factor because, as we will see in
Chapter 21, it must be greater than unity for macroscopic plasma stability in
a tokamak. Typical radial profiles for the poloidal and toroidal currents and
magnetic fields and the consequent q profile are shown in Fig. 3.6. As indicated,
q typically ranges from about unity on axis to a value of 35 at the plasma
edge. In terms of q the helical pitch of the field lines in (3.73) becomes simply
d/dz = 1/R0 q.

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

27

Figure 3.6: Radial profiles for a typical large aspect ratio tokamak: a) currents,
b) magnetic fields, and c) toroidal winding number q.
It is also customary in tokamaks to use the poloidal rather than toroidal
magnetic flux as the radial variable and to leave out the 2 factor by defining
pol
= R0

2

dr% B (r% ),

= R0 B (r)
er ,

poloidal flux function.


(3.75)

Thus, the normal magnetic flux representation of the screw pinch model for a
tokamak is
ez +
ez ,
B = (q z/R0 ) & B0

screw pinch field,

(3.76)

in which the approximate form indicates the lowest order form in the large
aspect ratio limit 0 << 1. Working out the magnetic field components from
either (3.66) or (3.76) using (3.74), we obtain
'
(
'
(
r
0

e & B0
e ,
ez + B
e = Bz (r)
ez +
ez +
(3.77)
B Bz
R0 q(r)
q
Note that the total magnetic field strength in this model is
3
3
B = Bz2 + B2 = Bz 1 + B2 /Bz2 = Bz h & B0

in which we have defined the geometric factor


3
8
h B/Bz = 1 + r2 /R02 q 2 = 1 + 02 /q 2 & 1.

(3.78)

(3.79)

For typical tokamaks 0/q 0.1 << 1, so usually the poloidal () magnetic field
is smaller than the toroidal (dominant) magnetic field by about an order of
magnitude. Thus, for typical tokamaks the approximate equalities at the end of
equations (3.69), and (3.76)(3.79) and subsequent ones in this section apply.
Note also that hence the helical field lines in typical tokamaks have only a slight
twist angle (torsion): r d/dz = r/R0 q = 0/q << 1.
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

28

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

In the screw pinch model the magnetic field strength is constant along a
magnetic field line; hence from (3.3) the screw pinch model magnetic field unit
B/B neither converge or diverge. However, the magnetic field in
vectors b
this model does have torsion, curvature, a perpendicular gradient and shear.
For the screw pinch model the unit vector along the magnetic field is
%
&
%
&
r
0
B = 1
ez +

e &
e .
ez +
(3.80)
b
B
h
R0 q
q
N = /|| =
Using the definition of the torsion in (??) with a unit normal

er , we find that for the screw pinch model

)(b
er ) =
= (b

e
1
(
e )
= 2

e ,
hR0 q
h
h R0 q

torsion.

(3.81)

Here, we have used the vector identities (??), (??) and (??) along with =

e /r to show that
e =
e (
e ) =
e (r) =
e (
er
e )/r =
er /r;
(
e )
(3.82)
e )
e = 1/r. Thus, the distance along a magnetic field line over
hence
er (
which it twists helically through one radian in the screw pinch model is
L = 1/r = h2 R0 q & R0 q,

torsion length.

(3.83)

The torsion vector can also be written in terms of the magnetic field comer see Problem 3.19. Note also that in the
ponents as = (B Bz /rB 2 )
tokamak limit of 0/q << 1 the helical pitch of the field lines given in (3.73)
becomes simply the torsion r .
The curvature in the screw pinch model is worked out similarly using the
vector identities (??) and (3.82):
)b
=
(b

r
r
r
(
e )

e =

er ,
hR0 q
hR0 q
(hR0 q)2

curvature. (3.84)

The curvature of magnetic field lines in the screw pinch model can be writer see
ten in terms of the magnetic field components as = (B2 /rB 2 )
Problem 3.19. The curvature length RC 1/|r | = (hR0 q)2 /r & R0 q(q/0) is
much longer than the torsion length L & R0 q in the screw pinch model of a
tokamak because the curvature is produced only by the poloidal motion of the
small pitch helical field lines. The perpendicular (radial) gradient scale length
[LB B/(dB/dr)] is of the order of the curvature radius RC . However, since
the difference depends on the current and plasma pressure profiles, it will not
be worked out until Chapter 20. Note also that since the curvature is only in
the radial direction there is only normal curvature. Because the magnetic field
lines do not have curvature within a magnetic flux surface, there is no geodesic
curvature see (??) in Section D.6.
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

29

Finally, it can be shown (see Problem 3.25) that for the screw pinch model
the local magnetic shear defined in (??) and (3.25) becomes (see Problem 3.19
for the form of the magnetic shear in terms of the magnetic field components
B and Bz ):
% &
1
r
d 1

er ) =
= 2
.
(3.85)
(b
er ) (b
LS
h R0 dr q
Note that in the screw pinch model the shear is constant on a magnetic flux surface (r = constant). Recalling from (3.74) that 1/q is just the radian rotational
transform /2 of the helical field lines, the local shear can be written as
=

r
s
s
1
r
d
dq
= 2
= 2
&
,
=
LS
2h2 R0 dr
h R0 q 2 dr
h R0 q
R0 q

magnetic shear,
(3.86)

in which
s(r)

r dq
,
q dr

magnetic shear parameter,

(3.87)

is an order unity magnetic shear parameter commonly used in stability analyses


of tokamak plasmas. There is magnetic shear in large aspect ratio tokamaks
only if.the axial current density Jz varies with radius r since B (r)/r
r
(1/r2 ) 0 r% dr% Jz (r% ). By convention, in tokamak plasma analyses the sign of
the shear is reversed so that s > 0 indicates positive or normal magnetic
shear, and s < 0 indicates reversed or abnormal shear.
Having delineated the local differential properties in the screw pinch model,
we can now develop a sheared slab model for it. At finite r since the curvature
and perpendicular gradient scale lengths are so long (compared to the torsion
and shear lengths) their effects are usually neglected in the simplest slab models.
As indicated previously, the sheared slab model does not include torsion effects.
Thus, the local sheared slab model for the screw pinch model of a tokamak near
a field line at a radius r0 is simply

+ b

e ].
(3.88)
Bss = B0 b
aux = B0 [ b + (x/LS )
in which
aux

x2
B0
,
2LS

and
e

%
&
%
&
1
0
0

b
er =

e
ez &
ez .
e
h
q
q
(3.89)

Here, the sheared slab model coordinates x, y, z correspond to r r0 , r0 [


(0/q)(z/R0 )], z + (0/q)r0 and we identify the directions in terms of the cylin
drical coordinate directions through the directions indicated in the unit vector b
within the r = constant
in (3.80) and a unit vector
e that is perpendicular to b
(magnetic flux) surface.
The preceding discussion focused on the screw pinch model for tokamaks.
The screw pinch model can also be used to represent reversed field pinch (RFP)
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

30

plasmas. In RFPs the toroidal and poloidal currents are much larger than those
in a tokamak (by a factor 1/0 A >> 1). In particular, the poloidal current
J is so large that it causes the toroidal magnetic field to reverse direction in the
edge of the plasma [see (3.69)] hence the name of the confinement concept.
The toroidal current in an RFP produces a poloidal magnetic field B that is
so large that q 0 and the small 0/q expansion that is used for tokamaks is
inappropriate. Such a large poloidal magnetic field also produces an order unity
helical pitch of the magnetic field lines; magnetic field lines in an RFP rotate
poloidally and toroidally on about the same length scales, and even become a
reversed direction helix (q < 0) in the edge of the plasma. For such a magnetic
field structure the curvature is clearly dominated by the poloidal motion of
the field lines; the toroidal curvature effects are higher order. Thus, to lowest
order the general [before the approximate equalities (&)] screw pinch model
developed in this section is often used to approximately describe reversed field
pinch plasmas. When a more precise description including toroidicity effects
is needed the full magnetic flux description developed in the following section
must be used.

3.6

Axisymmetric Toroidal Configurations*

For toroidal magnetic field plasma confinement systems with two magnetic field
components (toroidal, poloidal) a tremendous simplication occurs when the system is symmetric in the toroidal direction. Then, axisymmetric magnetic flux
surfaces are guaranteed to exist and both a Clebsch and flux surface representation are available. The resulting magnetic field system is the simplest, nontrivial
toroidal magnetic field system and is the basic paradigm for all types of toroidal
magnetic confinement systems.
In this section we develop the commonly used axisymmetric toroidal magnetic field descriptions and coordinate systems in general without using a
large aspect ratio expansion. We also show the relationship of the descriptions
and coordinates to the large aspect ratio tokamak and screw pinch models. At
the end of the section we discuss how the general axisymmetric toroidal model
can be used to describe other axisymmetric toroidal magnetic configurations.
The geometry we consider for an axisymmetric tokamak is shown in Fig. 3.7.
Since the toroidal magnetic field is in the direction of axisymmetry () and
=
e /R in which R is the major radius, it can be written as
e = R Btor I ,
Btor = Btor

toroidal magnetic field,

(3.90)

in which we have defined


I R Btor ,

toroidal field function.

(3.91)

Because of the axisymmetry, I must be independent of : I/ = 0. The


toroidal field function I can be related to the current flowing in the poloidal ()
direction. The poloidal curent flowing through a disk of (major) radius R that
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

31

Figure 3.7: Axismmetric tokamak coordinates (, , ) and geometry for calculating the poloidal current and magnetic flux.
is perpendicular to the axis of symmetry, as shown in Fig. 3.7, is given, using
Amperes magnetostatic law B = 0 J and Stokes theorem (??), by
Ipol

++

dS() J =

+
d" B/0 =

R d Btor /0

= (2/0 ) RBtor = (2/0 ) I.

(3.92)

Here, the minus sign occurs because the differential line element on the curve
C along the perimeter of the surface S is in the direction: d"
e d.
Thus, the toroidal field function I represents the poloidal current Ipol flowing
in the plasma and coils outside it. For isotropic pressure plasmas I = I(), i.e.,
I/ = 0.
In the limit of no current flowing in the plasma, the toroidal field function
I is constant and determined by the poloidal currents flowing in the toroidal
magnetic field coils around the plasma. Then, as can be inferred [see (??)] from
the magnetic field caused by current flowing in an infinite wire on the symmetry
axis (R = 0), the vacuum toroidal magnetic field strength decreases as one over

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

32

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS


the major radius R:
Btor = I0 /R = B0 R0 /R,

vacuum toroidal magnetic field strength,

(3.93)

in which B0 and R0 are the magnetic field strength and major radius at the
magnetic axis.
Next, we develop a form for the poloidal magnetic field Bpol . Using the
magnetic flux definition in (3.36) and taking account of the axisymmetry in the
toroidal () direction, the poloidal magnetic flux can be written in terms of the
e A = R A):
toroidal component of the vector potential (Ator
++
,
+ 2
dS() B =
d" A =
R d Ator = 2R Ator . (3.94)
pol =
S

For simplicity in the final tokamak magnetic field represention, it is convenient


to define a normalized poloidal magnetic flux function:
pol /2 = R Ator ,

poloidal flux function.

(3.95)

Since by definition / = 0, the poloidal flux fuction is independent of


the toroidal angle but in general depends on the cylindrical-like coordinates
in a = constant plane: = (r, ). In terms of this poloidal flux function
the toroidal component of the vector potential can be written Ator = /R,
e = . Thus, using (3.40), the magnetic
or vectorially as Ator = (/R)
field component produced by this magnetic flux becomes
Bpol = Ator = ,

poloidal magnetic field.

(3.96)

The strength of the poloidal magnetic field is


Bpol = || = ||/R,

poloidal magnetic field strength,

(3.97)

which shows that || = RBpol . The magnetic axis of the tokamak is defined
to be where Bpol = 0 and hence = 0.
Adding the two components of the magnetic field, the total magnetic field
becomes simply
B = Btor + Bpol = I + ,

axisymmetric magnetic field.


(3.98)

While this form is quite compact, it is unfortunately in neither a Clebsch form


nor a two component magnetic flux form. Also, it is not written in terms of
straight-field-line coordinates, and it is a mixed covariant and contravariant
form see Section D.8. Nonetheless, because this representation is compact
and rigorously valid it is heavily used in analyses of axisymmetric toroidal and
in particular tokamak plasmas.
Since by axisymmetry the poloidal flux function must be independent of the
toroidal angle (i.e., / = 0), taking the dot product of B with we obtain
B = (Btor + Bpol ) = I + = 0.
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

(3.99)

Fundamentals of Plasma Physics

33

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

Thus, the poloidal flux function satisfies the flux surface condition (3.37);
hence, magnetic field lines in axisymmetric toroidal systems lie on = constant surfaces and will be a convenient magnetic flux surface label and radial
coordinate. Note that, by construction and because of axisymmetry, is a suitable magnetic flux function for both the toroidal and poloidal magnetic fields.
Thus, we can develop a combined magnetic fluxes and Clebsch magnetic field
representation like (3.45) based on it.
So far we have identified two useful curvilinear coordinates for describing
the tokamak magnetic field: the axisymmetry angle for the toroidal angle and
the poloidal magnetic flux function for the radial variable. Next, we need
to identify a useful poloidal angle variable. We would like to have a poloidal
angle coordinate in which magnetic field lines are straight. Thus, we would like
a poloidal angle such that the magnetic field representation could be put in the
combined Clebsch and magnetic flux representation given by (3.45), with /2
replaced by 1/q and tor /2 replaced by for a tokamak representation.
In order to put the tokamak magnetic field (3.98) in the form of (3.45), the
toroidal magnetic field (3.90) must be put into the straight-field-line form
Btor = (q)

B = (q )

(3.100)

in which f (tokamak convention) is the desired straight field line poloidal


angle. Taking the dot product of the two forms of Btor given in (3.90) and
(3.100) with and equating them, we obtain
Btor = I = I/R2
= (q) = q () = q Bpol (3.101)

in which we have used q = 0 [because q is only a function of see


(3.105) below] and the order of the vector operations has been rearranged using
(??) and (??). Equating the results on the two lines of (3.101), we find
Bpol = I/qR2 = B ,

(3.102)

where the last equality follows from the fact that since by axisymmetry the
angle must be independent of , Btor = 0; thus, B = Bpol .
Defining a differential length d!pol in the poloidal direction on a magnetic flux
surface, the last form of (3.101) yields
+
1
I
1 !pol d!pol
I

=
=

=
,
(3.103)
!pol
B !pol qR2
q
B !pol R2

in which the integration is to be performed at constant , . The poloidal length


variable !pol can be defined in terms of the ordinary cylindrical angle about the
magnetic axis. Taking the dot product of the field line equation (3.28) with
and !pol , we find that the poloidal length variable is related to the cylindrical
angle by
B
dx
=
,
d!
B

DRAFT 22:52
September 22, 2003

dx !pol
B !pol
=
d!
B

c
!J.D
Callen,

d
d!pol
=
.
B
B !pol
(3.104)

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

34

Integrating the last form of (3.103) over one complete poloidal traversal of
the flux surface, which we define to be = 2, we obtain an expression for the
toroidal winding number:
,
,
d!pol
1
d
I
I
1
=
q() =
2
B !pol R2
2
B R2
or,

q() =

1
2

B
1
&
B
2

rBtor
rBtor
&
[1 + O{02 }].
RBpol
RBpol

(3.105)

Here, we have used I/R2 = B from (3.101) and indicated in the approximate equalities (here and below) the forms that result in the large aspect ratio
limit (0 << 1). Note that the lowest order, approximate form for q agrees with
the screw pinch model result (3.74).
The toroidal winding number q may be an integer or the ratio of two integers
(e.g., q = m/n); then, a magnetic field line on that surface would close on itself
after an integer number of poloidal (n) and toroidal (m) transits around the
torus see (3.74) and Fig. 3.8. Such a surface is called a rational surface.
All magnetic field lines on a rational surface rotate with the same rotational
transform , running forever parallel to adjacent field lines on the flux surface;
hence, they sample only a given field line on the flux surface. On the other hand,
if q is not the ratio of two integers, then the flux surface is called irrational. Thus,
we define (see Fig. 3.8)
q() = m/n,

rational flux surface,

q() -= m/n,

irrational flux surface,

(3.106)

in which m, n are integers. Magnetic field lines on an irrational flux surface


do not close on themselves; however, if they are followed long enough, they
fill the entire flux surface. This is called ergodic behavior since all points on
the surface are then equally sampled at least statistically in an asymptotic
limit. Note that the vast majority of flux surfaces are irrational; they form a
dense set. Rational surfaces are infrequent, separated radially (i.e., in ) and
of measure zero. Nonetheless, they are very important in magnetized toroidal
plasmas because physical processes taking place on adjacent rational field lines
are mostly isolated from each other, and because they are degenerate field lines
that are especially vulnerable to resonant nonaxisymmetric perturbations that
can produce magnetic island structures like those discussed in Section 3.3.
An explicit expression for will now be obtained. Using the rigorous form
of the definition of q in (3.105) and defining like the simple geometric angle
to be zero on the outer midplane of the torus, we can develop from (3.104)
explicit expressions for the straight-field-line poloidal angle:
+
+
+
I
I
1 d
1
B
1 ! d!
. (3.107)
=
=
d

q 0 B ! R2
q 0 B R2
q 0
B
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

35

Figure 3.8: Puncture plots of magnetic field lines in a = constant plane on


a magnetic flux surface in an axisymmetric torus. The numbers listed indicate
the number of toroidal transits executed by a field line. On rational surfaces
field lines retrace the same trajectory after m toroidal transits whereas on an
irrational surface a single field line is ergodic and (eventually) samples the entire
surface.
It can be shown using steps like the last few ones in (3.105) that, to lowest order
in a large aspect ratio expansion (e.g., near the magnetic axis), the straight field
line coordinate is equal to the local cylindrical coordinate :
= O{0 sin }.

(3.108)

The order 0 sinusoidal variations of with depend on the currents flowing


in the plasma; their evaluation will be deferred until Chapter 20 where we use
the macroscopic force balance equations in a finite-pressure tokamak plasma to
determine the currents in a tokamak plasma and the shape of the (x) surfaces.
As can be seen from Fig. 3.1d, the magnetic field in a tokamak has parallel
and perpendicular gradients, curvature (both normal and geodesic), and local
torsion and shear that are not constant along the magnetic field. Below, we will
give general expressions for each of these properties both in general, and also in
their lowest order forms in a large aspect ratio (0 << 1), low plasma pressure
expansion. To lowest (zeroth) order the magnetic flux surfaces become circles
about the magnetic axis. (To first order in 0 the flux surfaces are still circles, but
their centers are shifted outward slightly in major radius see Section 20.4.)
Thus, to lowest order we will use the r, , z coordinates of a cylinder whose z
axis lies on the magnetic axis of the tokamak, of a type shown in Fig. 3.1d. To
lowest order the model will mostly reduce to the screw pinch model discussed
in the preceding section compare Fig. 3.5 with Fig. 3.7.
The major radius R to any point in the plasma will be given in terms of the

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

36

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

major radius of the magnetic axis (R0 ) and the local cylindrical coordinates by
R = R(, ) & R0 [ 1 + 0 cos + O{02 }].

(3.109)

Using this approximate representation in the equation for the vacuum magnetic
field strength variation with R given in (3.93) and the fact that in our tokamak
model B & Btor [ 1 + O{02 /q 2 }], we obtain
B = B(, ) & B(r, ) & B0 [ 1 0 cos + O{02 }], tokamak field strength.

(3.110)

The magnetic field strength in a tokamak varies approximately sinusoidally


along a helical magnetic field line from its minimum on the outside ( = 0 where
Bmin /B0 & 1 0) to its maximum on the inside ( = where Bmax /B0 & 1 + 0)
of the torus. Thus, it can be represented by the sinusoidal model in (3.6) using
! R0 q and L! 2R0 q. The magnitude of the variation along a magnetic
field line is usually small: B Bmax Bmin & 20B0 << B0 . Hence, the magnetic mirror ratio defined in (3.7) is usually only slightly greater than unity:
Rm Bmax /Bmin & 1 + 20. In summary, the variation of the magnetic field
strength along field lines in large aspect ratio tokamaks can be modeled by (3.6)
with
Bmin & (1 0)B0 ,

B & 20B0 , ! R0 q , L! 2R0 q,


tokamak Bsin model parameters.

(3.111)

To calculate the perpendicular gradient, curvature and shear in the tokamak


magnetic field we need to explicitly relate the tokamak magnetic flux system
coordinates , , , which are unfortunately not orthogonal ( -= 0), to
the local cylindrical coordinates (r, , z) about the magnetic axis:
+ r
er , (r) &
dr Bpol R0 ,
(3.112)
& Bpol R r & Bpol R0
0

I
B0
Bpol

e
, B =
&
,
&
& &
r
qR2
r
R0 q

e
&
=
[ 1 0 cos + O{02 }].
R
R0

(3.113)
(3.114)

The poloidal magnetic field strength oscillates slightly with poloidal angle:
Bpol ||/R & (0/q)B0 [ 1 + O{0 cos }].

(3.115)

In calculating gradients of various quantities in tokamak system coordinates,


we just use chain rule differentiation:
%
&
1
B
1
B
B
=

+
&
[
er cos +
e sin + O{0}] .
B
B

R0
(3.116)
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

37

in which we have used (3.93) and / terms vanish by axisymmetry. In a tokamak with low plasma pressure and hence a small poloidal current (see Chapter
20) the JB contribution to the curvature is small. Thus, using (3.20) the
curvature is given by
%
&
1
B
B
JB
B
+ 0
&

+
=
B
B2
B

1
[
er cos +
e sin + O{0}] , tokamak curvature vector. (3.117)
&
R0
To this lowest order the curvature is simply the toroidal curvature of the system.
Note that a tokamak has both normal (perpendicular to the flux surface, N &
r = cos /R0 ) and geodesic (within the flux surface, B & = sin /R0 )
curvature see Section D.6. Note further that, because of the inclusion of
toroidicity effects, the tokamak curvature is one order in 0 larger than that in
the screw pinch model (3.84); however, its sign oscillates with the poloidal angle
and its average is of the same order as the curvature in the screw pinch model.
To determine the O{0} terms in (3.117) we need to take account of the plasma
pressure and current profiles in the tokamak see Chapter 20.
Using a number of vector identities and other manipulations (see Problem
3.29), it can be shown that the normal torsion in a tokamak can be written as
'
%
&
%
&(
I ||

qR||

qR||
(B
)
+
N =
B2R

1
[ 1 + O{0 cos }], tokamak local torsion,
(3.118)
&
R0 q
which to the lowest order is the same as in the screw pinch model see (3.81).
Similarly, the magnetic shear in a tokamak becomes (see Problem 3.31):
'
'
(
%
&(
||2

||2
dq

(B
)
(q)
=

(B
)
+
q
y =
B2

B2
d

1 + O(0 cos )
[ s O{0 cos }], tokamak local shear,
(3.119)
&
R0 q
which again to lowest order is the same as in the screw pinch model see
(3.86). [The convention in the tokamak literature is to reverse the sign of the
shear so that it is positive for normal tokamaks in which q increases with
radius (see Fig. 3.6).] Note that both the local torsion and local shear have order
0 sinsoidal variations along a magnetic field line as it moves from the outside to
inside of the torus but their averages over a magnetic flux surface (i.e., over
) are approximately given by their respective values in the screw pinch model.
Again, to obtain the next order (0) terms correctly we need to take account of
plasma currents and pressures, which we defer until Chapter 20.
In this section we have developed the magnetic field representation and properties of axisymmetric toroidal magnetic field systems in general, and then indicated the lowest order results in the large aspect ratio expansion (0 r/R0 =
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

38

1/A << 1) after an approximate equality sign (&). While the discussion has focused on the tokamak magnetic field structure, the general development applies
to any axisymmetric toroidal system. Thus, it applies to spherical tokamaks
[very low aspect ratio (A 1.11.5) tokamaks], spheromaks [effectively unity
aspect ratio tokamaks without toroidal field coils] and reversed field pinches
[effectively tokamaks with very low q 0 << 1]. For spherical tokamaks and
spheromaks the full generality of the magnetic flux coordinates must be used
because a large aspect ratio expansion is invalid except for small radius flux
surfaces near the magnetic axis where a large aspect ratio expansion can be
used. As indicated at the end of the preceding section, the general screw pinch
model represents reversed field pinches except for the purely toroidal effects
(variation of B from the outer to inner edge of the torus and toroidal curvature of field lines). These latter effects are included in the general magnetic
flux model developed in this section. In summary, the general magnetic flux
surface model developed in this section is appropriate for describing all types of
axisymmetric toroidal magnetic field configurations tokamaks, reversed field
pinches, spherical tokamaks, and spheromaks.

3.7

Local Expansion of a Magnetic Field+

In order to develop a comprehensive picture of all the possible first derivative


properties of a magnetic field, in this section we carry out a formal Taylor series
expansion of the magnetic induction vector field B(x). The expansion will be
carried out at an arbitrary point in the magnetic configuration where the origin
of a local Cartesian coordinate system will be placed. Thus, our local Taylor
series expansion (subscript le) becomes
Ble (x) = B0 + x B0 + . . . ,

(3.120)

in which B0 is the magnetic induction field at the chosen point, x is the vector
distance from this point, and B0 represents the evaluation of the tensor B at
this point. The second and higher order terms in (3.120) will be neglected since
we are interested here only in the local properties of non-pathological magnetic
fields for which the first derivatives provide a sufficient description.
This section uses a number of vector differentiation identities and seeks to
connect the local magnetic field derivatives to the common definitions of most
of these properties for arbitrary vector fields. These subjects are summarized
briefly in Appendix D, and in particular in the Section D.6. Readers are encouraged to read the relevant sections in Appendix D in conjunction with this
section.
As usual we decompose the magnetic induction field B(x) into its vector
direction and scalar magnitude components at any spatial point x by writing
= B(x) b(x),

B(x) = B b

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

(3.121)

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

39

in which

B(x) |B(x)| = B B, magnetic field strength,

b(x)
B(x)/B(x) = B/B, unit vector along B(x).

(3.122)
(3.123)

Using this decomposition, we can write the tensor B0 as


0 = b
0 B0 + B0 b
0 ,
B0 (B b)

(3.124)

0 is the unit vector b


evaluated at the origin (point of interest) and
in which b

B0 , b0 are B, b evaluated at this same point.


and hence B tensors we
To work out the various components of the b

can use a local Cartesian coordinate system whose z axis is aligned along b

at x = (0, 0, 0) (i.e., b0 ) and which has its x axis in a particular direction


0 , which we will specify later. Thus, the orthonormal triad
perpendicular to b
0 ,
of unit vectors characterizing this local coordinate system will be
ez = b
ex ,

ex . For notational simplicity, in what follows we omit the subscript

ey b0

zero on the magnetic field unit vector b.


Consider first the components locally parallel to the magnetic field:
zz
(b)
zx
(b)
zy
(b)

(b)xz
yz
(b)

(b
)b
= b
b(

= 0,
b
b)
)b]

)b
-= 0,
[(b
ex =
ex (b
)b]

)b
-= 0,
[(b
ey =
ey (b
1

= 0,
b (
ex )b = 2 (
ex )(b b)
1

= 0,
b (
ey )b = (
ey )(b b)

(3.125)

is a unit vector, we have b


b
= 1,
in which we have used the fact that since b
and the vector identity,
1
b)
= 0 = b(

+ (b
)b.

(b
b)
2

(3.126)

For the components locally perpendicular to the magnetic field we have


xx
(b)
yy
(b)

(b)xy
yx
(b)

ex (
ex )b,

ey (
ey )b,

ey (
ex )b = +
ey
ey
ex (b
ey ,

ex (
ey )b =
ex
ex +
ex (b )
ey ,

(3.127)

in which we have worked out the last two cross terms using vector differentiation
identities (??) and (??) as follows:
=
+
b
(
)
ex )b
ey [(
ex b)
ex ( b)
ex ) + (b
ex ]

ey (
)
)
ey +
ey (b
ex =
ey
ey
ex (b
ey ,
=
ey
(3.128)
yx .
and similarly for (b)
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

40

Taking the derivative of the equation along a magnetic field line given in
(3.28) yields the second derivative of the field lines coordinate and hence the
)]:
local curvature of the magnetic field [d/d! (b

d2 b(!)
)b
= b(

= (b
b),
d!2

curvature vector,

(3.129)

= 0). The zx, zy


which is perpendicular to the local magnetic field (b

components of the tensors b can be written in terms of the x, y components


of the curvature vector:
)b
=
zx =
ex (b
ex x ,
(b)
)b
=
zy =
ey (b
ey y .
(b)

(3.130)

The x component of the curvature vector is the same as the x in the sheared
slab model given by (3.16) and (3.21). From geometrical considerations the
radius of the curvature vector RC is antiparallel to the curvature vector .
Hence it is given by
RC = /2 ,

2
or = RC /RC
,

radius of curvature.

(3.131)

The torsion of a vector field is by convention defined to be the negative of the


ex =
ey :
parallel derivative of the binormal vector, which for our geometry is b
)
) (b
ex ) = (b
ey ,
= (b

torsion vector.

(3.132)

The x component of this vector is given by


)
)
ex (b
ey =
ey (b
ex ,
x

(3.133)

which is a quantity that appears in xy and yx components of the tensor b.


Physically, the torsion vector measures the change in direction (or twist) of the
ex as one moves along the magnetic field.
binormal b
Shear of a vector field can be defined for the two directions perpendicular to
the magnetic field by
ey ) (b
ey ),
ex
ex = (b
x

shear in surface perpendicular to x,


(3.134)

ex ) (b
ex ),
ey
ey = (b
y

shear in surface perpendicular to y.


(3.135)

That these quantities represent the local shear in the vector field B can be
ex represents the surface locally perseen by realizing that, for example, b
ex ) represents the tangential motion (see Fig. ??) or
pendicular to
ey , (b
ex ) (b
ex ) is the component
differential twisting of this surface, and (b
ex (or
ey )
of this differential twisting in the original direction
ey . Note that if
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

41

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

were a unit vector corresponding to a contravariant base vector ui in a curvilinear coordinate system (see Section D.8), then
ei = ui /|ui | and hence i
i
i
i 2
ei = (u u )/|u | = 0. Thus, there is no shear in a direction

ei
described by the gradient of a scalar function (e.g., a magnetic flux function)
because the gradient of a scalar is an irrotational quantity (f = 0). The
b
(or parallel component of the vorticity
corresponding parallel shear b
in the vector field B) can be written in terms of the x component of the torsion
vector and the two perpendicular shear components as follows:
b
= 2 x + x + y ,
z b

parallel component of vorticity in B field.


(3.136)

In the absence of shear, this relation is analogous to the component of rotation


of a rigidly rotating fluid in its direction of flow, i.e., 12 VV.
tensor can be written in
The xy and yx or cross components of the b
terms of the x, y components of the torsion and shear as follows:
xy = y + x ,
(b)
yx = x + x .
(b)

(3.137)

can be written as [using (??)]


The divergence of the unit vector b
0

9
:;
<

ex )b +
ey (
ey )b + b (b )b .
b =
ex (

(3.138)

= (1/B)[
ex (
ex )b
ex (
ex )B],
x

ey (
ey )b = (1/B)[
ey (
ey )B].
y

(3.139)

tensor
Thus, the xx and yy (or diagonal matrix element) components of the b
We
represent the x and y components of the divergence of the unit vector b.
define these divergence () components of the vector field B as follows:

Collecting together the various components


in matrix form,

x
x + y
x x
= (
y
ey b)
b
ex
x
y

we thus find,
of the tensor b

ex
0
ey .
0

0
b

(3.140)

Further, using this result in (3.124), we find that the tensor B can be similarly
written as

ex
Bx /x By /x Bz /x
1
0
Bx /y By /y Bz /y
ey
ey b
B
ex

Bx /z By /z Bz /z
b
(3.141)

ex
x
x + y x
0
1
x x
ey ,
y
y
ex
ey b
= B0

x
y
z
b
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

42

in which the differential parameters of the magnetic field, including some new
gradient () ones which we have introduced here, are defined by:
ex (
ex )B], y B1 [
ey (
ey )B],
x B1 [
1
divergence,
z B (b )B,

ex (b )b,
y
ey (b )b,
curvature,
x
)
)
x
ex (b
ey ,
y
ey (b
ex = x , torsion,
ex
ex ,
y
ey
ey ,
shear,
x
ex ln B,
y
ey ln B,
x
ln B0 = z ,
gradient B.
z b

(3.142)

Using the expression for B in (3.120) yields the following Taylor series expansion for the magnetic induction field B:
( 1 + x ln B)
Ble & B0 [ b
+ z
+ x (x
ey y
ex )
ey y x
ex )
+ (x y
ex + y y
ey ) ]
+ (x x

lowest order + gradient (B)


)b
= x
ex + y
ey
curvature, (b
2

torsion, x
ey y
ex = 12 b(x
+ y2 )
shear
divergence
(3.143)

( 1 + xx + yy + zz )
= B0 [ b
ey (zy + xx + xy + yy )].
+
ex (zx yx yx + xx ) +
Note that this general result simplifies to the sheared slab model (3.8) when the
parameters y , z , y x , x , x , y and z all vanish, i.e., when the magnetic
field does not vary in the y, z directions, and there is no shear in the x direction
and no torsion of the magnetic field lines.
The solenoidal condition ( B = 0) will be satisfied by this local expansion
as long as
0=

1
1
1
ln B = x + y + z .
Ble = [
ex (
ey (
ex )B] + [
ey )B] + b
B
B
B
(3.144)

Thus, the three diagonal components of the matrix of B tensor elements are
not independent; there are only 8 independent components of the B tensor.
The curl of our local approximation of the magnetic induction field B is
given by

(b
b)
],
ln B + ) + b
Ble = B0 [ b(

(3.145)

in which we have made use of (3.136). As in the sheared slab model, for plasma
equilibrium situations where the magnetostatic Amperes law applies, the currents flowing in the plasma provide further constraint relations between the
various local differential parameters involved in (3.145). In particular, since currents perpendicular to magnetic fields are typically small, usually & ln B
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

43

b(
ln B), i.e., the curvature vector is approximately equal to the
perpendicular gradient of the magnetic field strength. Also, the parallel shear
b
= 0 B J/B 2 ) is
(parallel component of vorticity in the magnetic field b
nonzero only if current flows along the magnetic field, i.e, B J -= 0.
The curvature, shear and perpendicular gradient properties of a magnetic
field were discussed in the context of the sheared slab model in Section 3.1.
They are illustrated in Fig. 3.2 and their effects mathematically described in
(3.14) and (3.24). The additional magnetic field line properties of torsion and
divergence can be understood as follows.
Eliminating all but the torsion terms in (3.143), the x, y, z equations governing the trajectory of a magnetic field line deduced from (3.28) become simply
dy
= x x ,
d!

dx
= y x ,
d!

dz
= 1 = ! = z.
d!

Dividing the second equation by the first yields


x
dy
=
dx
y

d (x2 + y 2 ) = 0,

whose solution is
x2 + y 2 = x20 + r02 r02 = constant.
This result can be used to reduce the equation for dx/d! = dx/dz to one in only
two variables:
dx
= y x = (r02 x2 )1/2 x
dz

arcsin

x
= z x + constant.
r0

The equations governing a field line with torsion x that passes through the
point x0 , y0 are thus given by
x = r0 sin(zx + 0 ),

y = r0 cos(zx + 0 ),

0 arctan(y0 /x0 ).
(3.146)

These equations show that torsion means that as one moves along a magnetic
field line it undergoes circular motion through an angle of one radian in an axial
distance of x1 L in the plane perpendicular to the magnetic field see
Figure 3.9a. Comparing the field line trajectory equations for torsion with that
derived previously for shear, (3.24), or Fig. 3.9a with 3.9b, we see that whereas
torsion represents rigid body rotation or twisting of the field lines in the
plane perpendicular to the magnetic field, shear (x = 0, y -= 0 for Fig. 3.2b)
represents differential twisting of field lines out of a plane (the xz plane for
y -= 0). Thus, whereas the torsion terms can be removed by transforming to a
rotating coordinate system, the effects of magnetic shear cannot be removed by
such coordinate transformations.
To explore the divergence of magnetic field lines we eliminate all but the x
and z terms in (3.143), and take z = x so as to satisfy the solenoidal condition (3.144). Then, the equations governing the magnetic field line trajectory
DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

44

Figure 3.9: Additional magnetic field line characteristics. In each sketch the
nature of magnetic field lines are indicated when only the identified coefficient(s)
do not vanish.
in the xz plane become
dx
= x x ,
d!

dz
& 1.
d!

Integrating the first equation a short distance (z x << 1) along the magnetic
field line that passes through the point x = (x0 , 0, 0) yields
'
(
x0 B(0)
d ln B
x z
z z
&
.
(3.147)
x(z) = x0 e
= x0 e
= x0 exp z
dz
B(z)
This result shows that the divergence (x > 0) in the xz plane (cf., Fig. 3.9b)
is accompanied by a decrease in the magnetic field strength [z d ln B/dz < 0,
B(z) < B(0) for z > 0], as is required by the solenoidal condition B = 0
see (3.2) and (3.144). The divergence scale length x1 (or equivalently |z |1 )
is the linear extrapolation distance along a field line over which the density of
magnetic field lines would decrease (increase for x < 0, z > 0) in magnitude
by a factor of two and the field lines diverge (converge) by a factor of e.
We can also develop a Taylor series expansion of the magnetic induction field
bx)

as the
B(x) about a given magnetic field line. Here, we define x = b(
small vectorial distance off a given magnetic field line in the x, y plane locally
perpendicular to the field line. The axial distance along the magnetic field line
is parameterized by the length ! along it from an initial reference point. Thus,

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

45

the desired expansion is


0 (!) + x B0 (!)
Ble (x , !) = B0 (!) b
0
0 (!) [ 1 + x ln B0 (!) ]
= B0 (!) b

1
0 (!)(x2 + y 2 ) + [x y
+ 12 x b
ey y x
ex ] + [x x
ex + y y
ey ]

(3.148)

in which all quantities are now evaluated on a particular field line (i.e., at
x = 0), but the functional dependence on ! remains. Compared to the expansion about a point given in (3.143), we see that the curvature () and parallel
gradient (z ) or divergence (z ) terms are missing from the expansion about a
0 (!) and B0 (!). Note also
field line because these effects are included via b
that this field line expansion for B satisfies the solenoidal condition B = 0 as
long as (3.144) is satisfied, and yields the result given in (3.145) for B. Since
charged particles in a plasma usually move much more easily (and hence traverse
much longer distances) along magnetic field lines than perpendicular to them,
the expansion about a field line is usually more useful, at least conceptually, for
plasma physics applications.
As discussed in Sections 3.2, 3.43.6, often in plasma physics there exist
a set of nested magnetic flux surfaces (x) that surround nested bundles of
magnetic field lines. When such surfaces exist, is locally perpendicular to
the magnetic field and it can be used to specify the directions of the unit vectors
in the plane locally perpendicular to the magnetic field:

ex /||,

ey b/|
b|
= b/||.

(3.149)

For such cases it is customary to call the curvature component in the direction
perpendicular to the magnetic flux surfaces (i.e., in the direction) the normal
curvature and the curvature component within the magnetic flux surface (i.e.,

in the b
direction) the geodesic curvature:
x = /||,

normal () curvature,

geodesic (b)
curvature.
y = b b/||,

(3.150)

Since this x coordinate direction is in the direction of the gradient of a scalar,


the shear in the x direction vanishes: x = /||2 = 0. However,
there can still be shear in the y direction; it can be written as
y =

(b)
(b)
,
2

|b|

shear for
ex .

)(b)

(b
,
2
||

torsion for
ex .

(3.151)

The torsion for this situation where magnetic flux surfaces are assumed to exist
can be written as
x =

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

(3.152)

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

46

Using these relations in (??) along with the fact that for
ex we have
ex
ex = (/||) (/||) = 0, we can write the parallel
x
shear in the magnetic field as
b
= 0 B J/B 2 = y + 2 x ,
b

parallel shear for


ex .
(3.153)

Note that this relation provides a relationship between the parallel current and
the torsion and shear in the magnetic field. It is the parallel current analogy
to the relationship between the curvature vector, ln B and perpendicular
current given in (3.20). Finally, the x and y field line divergence parameters
can be written as
(
ex ) (||1 ) = (b
) ln ||, (3.154)
ex )
ex = (b
x = b
) ln || (b
) ln B,
y = x z = (b

divergences with
ex .
(3.155)

These simplified formulas for the situation where the x coordinate is taken to
be in the magnetic flux surface gradient direction are the most commonly used
ones in plasma physics.
REFERENCES AND SUGGESTED READING
Plasma physics books that discuss some aspects of the structure of magnetic fields
include
Bittencourt, Fundamentals of Plasma Physics, Chapt. 3 (1986) [?]
Hazeltine and Meiss, Plasma Confinement, Chapt. 3 (1992) [?]
Miyamoto, Plasma Physics for Nuclear Fusion, Chapt. 2 (1980) [?]
Schmidt, Physics of High Temperature Plasmas (1966,79) [?]
White, Theory of Tokamak Plasmas, Chapts. 12 (1989) [?]
Comprehensive treatments of the structure of magnetic fields for plasma physics applications are given in
Morozov and Solovev, The Geometry of the Magnetic Field in Reviews of
Plasma Physics, M.A. Leontovich Ed., Vol.2, p.1 (1966). [?]
Dhaeseleer, Hitchon, Callen and Shohet, Flux Coordinates and Magnetic Field
Structure (1991) [?]

PROBLEMS
3.1 The magnetic field strength inside a solenoidal magnet composed of a series of
circular coils can be characterized by a uniform magnetic field B = B0
ez plus
sin(2z/L) in
a small ripple field whose magnitude on axis is given by B = B
<< B. Develop a
which L is the axial distance between the magnets and B
sinusoidal model of the type given by (3.6) for this situation; that is, specify
all the parameters of the sinusoidal model for this bumpy cylinder magnetic
field. Sketch the behavior of the field lines inside the solenoid using (3.3). /

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

47

3.2 First, show that the magnetic field around a wire carrying a current I in the z
direction is given by B = [0 I/(2r)]
e , where r is the radius from the center
of the wire. Next, show that the curvature vector for this magnetic field is
=
er /r and hence that the radius of curvature of the magnetic field lines is
r. Finally, show that for the simple magnetic field = B. /

3.3 Show by direct calculation starting from (3.15) that for a vacuum (J = 0)
magnetic field which can be represented by B = M one obtains =
ln B, as follows from (3.20) for this situation. //

3.4 Integrate the field line equation dx/dz = Bx /Bz for the sheared slab model to
obtain the field line trajectory in the y = constant plane. Why is the result
slightly different from that in (3.14)? /

3.5 Use (3.31) to determine a potential representation for the bumpy cylinder mag M (r, z). Check
netic field given in Problem 3.1 in the form of M = M 0 (z) +
your result by calculating |B| for your model and comparing it to the desired
result. //
3.6 Propose a suitable magnetic flux and a vector potential for a cylindrical model
of an infinite, homogeneous magnetic field, and show that they yield the desired
magnetic field. /
3.7 Calculate Clebsch , and % coordinates for the sheared slab model as follows.
First, write down three independent field line equations for dx/dz, dy/dz, and
dz/d%. Integrate the first two of these equations to obtain
A
B
(x) = B0 x + x2 /2LB z 2 /2RC = constant = x = x(z, )
+ z
dz " x(z " , )
1
(x) = y
= constant = y = y(z, , ).
LS 0 1 + x(z " , )/LB
Show that the indicated field line equations reduce to (3.14) and (3.24) near the
origin. In which directions do and point? Also, show that
yields the slab model field given in (3.8). Next, integrate the third field line
equation to obtain an expression for % that is correct through first order. Calculate %; in what direction does it point? Finally, calculate B %; explain why
your result is (or is not) physically reasonable. //
3.8 For the sheared slab model, why is the field line equation in (3.14) different from
that implied by constancy of the z in (3.41) or the given in the preceding
problem? /
3.9 Show that the toroidal and poloidal magnetic field components given in (3.44)
give the respective toroidal and poloidal magnetic fluxes. //
3.10 Show that when closed toroidal magnetic flux surfaces exist .the toroidal and
poloidal magnetic
fluxes can be calculated from 2tor = d3x B and
. 3
2pol = d x B in which the volume integrals are taken over a closed
toroidal flux surface. [Hint: A relevant volume for the toroidal surface tor
encloses a torus defined by surfaces that satisfy B = 0, but has a cut at
a = constant plane.] ///
3.11 Obtain the radial curvature R from a definition like (3.13) for an axisymmetric
magnetic mirror. /*

3.12 Develop a magnetic flux representation in the form (R, ) = 0 (R) + (R,
Z)
for the bumpy cylinder magnetic field given in Problem 3.1. Check your result
by calculating |B| for your model and comparing it to the desired result. //*

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

48

3.13 Give an approximate equation for the variation of the radius of a field line in
<< B0 . /*
the bumpy cylinder magnetic field given in Problem 3.1 when B
3.14 Write down the relevant field line equation for a dipole magnetic field. Integrate
this equation to determine r = r() along a magnetic field line in terms of the
radius r0 at the equatorial plane. Substitute your result into (3.60) to obtain
the variation of the magnetic field strength with along field lines. What is the
magnetic mirror ratio to a latitude of 45o ? Finally, show that near the equatorial
plane the magnetic
field strength can be represented by the quadratic well model

with L# = ( 2/3)r0 . //*


3.15 Consider a screw pinch model situation where current only flows parallel to the
with
magnetic field. Assume the parallel current density is given by J = J# (r)b
J# (r) 0 for 0 r a. What are the axial and poloidal current densities
for such a situation? Develop an expression for the toroidal magnetic field
strength from (3.69) for such a situation. Show that the poloidal current effect
on Bz in the large aspect ratio tokamak limit where */q 0.1 << 1 is of
order (*/q)2 102 << 1. Does the flowing current produce a diamagnetic or
paramagnetic effect in the region where the current J# is flowing? that is,
does it decrease or increase Bz inside the plasma? Finally, give an expression
for the radial variation of the total magnetic field strength for such a case. //*
3.16 Show for the screw pinch model of a large aspect ratio tokamak with Bz (r) ' B0
which has a well-behaved current density profile near the magnetic axis (i.e.,
dJz /dr = 0 at r = 0 so that Jz ' J0 + r2 J0"" /2), that q increases with radius and
can be approximated by q(r) = q(0)/(1 r2 /rJ2 ) near the magnetic axis. What
is the sign of J0"" for a profile peaked at r = 0? Determine expressions for q(0)
and rJ in terms of J0 and J0"" . ///*
3.17 The value of q usually decreases with radius away from the magnetic axis in a
reversed field pinch. Use the combination of the two preceding models to obtain
the necessary conditions on the current profile for this to occur. ///*
3.18 Consider a box axial current proflile given by Jz (r) = J0 H(r0 r) in which
H(x) is the Heaviside step function defined in (??). Calculate and sketch the
q(r) profile for this current profile in a screw pinch model of a tokamak for
0 r a 2 r0 . Why does q increase as r2 outside the current-carrying region
(i.e., for r0 r)? /*

3.19 Determine the forms of the magnetic field curvature vector , torsion vector
and local shear for the screw pinch model in terms of the magnetic field
B/B = (Bz
components using b
ez + B
e )/B in the appropriate definitions
of these properties of a magnetic field. Show that the results can be written as
%
&
%
&
B2
B Bz
Bz d B
B d rBz
= 2
er , =

e
,

.
r
rB
rB 2
B dr B
rB dr
B
Also, show that these results reduce to the forms given in (3.84), (3.81) and
(3.85), respectively. //*
3.20 In a reversed field pinch (RFP) the value of q(r) vanishes at the reversal surface
rrev )= 0 and it might seem from (3.86) and (3.87) that the magnetic shear is
undefined there. For the screw pinch model show that the magnetic shear can
be written as
%
&
r/R0
1
dq
shear =
=
.
LS
q 2 + r2 /R02 dr

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

49

What is magnetic shear length LS where q = 0? Evaluate this formula for LS


at the reversal surface for a model profile q(r) = q0 (q0 qedge ) r2 /a2 with
q0 = 0.2, qedge = 0.1125 for an RFP with R0 = 1.6 m and a = 0.5 m. //*

3.21 For tokamak plasmas with noncircular cross-sections (in the = constant plane)
it is customary to define an effective cylindrical winding number or safety factor:
qcyl 2Apol Btor0 /(0 R0 Itor ) in which Apol is the cross-sectional area and Itor is
the toroidal current through that area. Show that in the circular cross-section,
large aspect ratio tokamak limit qcyl reduces to the q given in (3.74). /*
3.22 Use projections of the field line definition (3.28) to obtain equations for d%/dz
and d%/d along a field line in the screw pinch model. Integrate the field line
equations to determine expressions for the length of a field line in terms of the
axial distance z and the poloidal angle traversed by the field line: % = %(r, z)
and % = %(r, ). //*
3.23 Attempt to obtain a sheared slab model around a radius r = r0 for the screw
pinch model by expanding the screw pinch model magnetic field in (3.77) taking
in (3.80) evaluated at r = r0 .
the slab model
ez to be in the direction of the b
Explain why the shear parameter obtained this way is different from the =
1/LS indicated in the sheared slab model given in (3.86). //*
3.24 Show that for the screw pinch model magnetic field
%
&
%
&
%
&
1 f
f
1
f
f
)f = 1 B f + Bz f = 1
(b
+
'
+q
B
r
z
h R0 q
z
R0 q

in which f = f (x) is any differentiable scalar function of space. /*


3.25 Give the steps used in obtaining the shear for the screw pinch given in (3.85).
er ) (b
er ) using the vector
[Hint: First show from (3.25) that = (b
er = (r/h) (r/hR0 q)z and use vector
identity (??). Then, show that b
identities (??), (??) and (??) in evaluating the shear .] //*
3.26 Use the magnetostatic Amperes law and Stokes theorem to obtain an expression
for the toroidal current flowing inside of a magnetic flux suface in terms of an
integral of the poloidal magnetic field in the axisymmetric toroidal model. Take
the large aspect ratio tokamak limit of your result and compare it to (3.68). /*
3.27 Use projections of the field line definition (3.28) to obtain equations for d%/d
and d%/d along a field line in the axisymmetric toroidal model. Show that
the ratio of these equations gives the tokamak field line equation d = q() d.
Integrate the field line equations to determine general expressions for the length
of a field line in terms of the toroidal and poloidal angles and traversed by
the field line: % = %(, ) and % = %(, ). Show that in the large aspect ratio
tokamak expansion % ' R0 = R0 q() and indicate the order of the lowest
order corrections to these results. ///*
3.28 Show that for the axisymmetric toroidal magnetic field given in (3.27)
%
&
%
&
f
f
f = B f + I f = I
b
+
q
B
BR2
BR2 q

in which f = f (x) is any differentiable function of space. [Hint: Use chain


rule rule differentiation to write = ()/ + ()/ + ()/.]
Also, show that in the large aspect tokamak expansion this result reduces to the
similar screw pinch limit result obtained in Problem 3.24. //*

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 3. STRUCTURE OF MAGNETIC FIELDS

50

3.29 Work out the general expression for the torsion for an axisymmetric toroidal
configuration given in (3.118). [Hint: First show that for the axisymmetric
toroidal magnetic field the normal torsion can be written in the form N =
)
(I/B 2 ||2 ) [() ()] using the first form of (b
from the preceding problem, and (??). Next, show using (??) and (??) that
N = (I/B 2 R2 ) (
ey
ey ) = (I/B 2 R2 ) (
ey
ey ) in which
ey
/|| = (qR||/I)[ ()( )/||2 ]. Finally,
work out the last form of N using (??), (??) and (??).] ///*
3.30 Show that the large aspect ratio tokamak expansion of the general expression
for local torsion yields the result indicated after the ' in (3.118). //*

3.31 Work out the general expression for the magnetic shear in an axisymmetric
toroidal configuration given in (3.119). [Hint: Use the form of B in (3.100) and
work out B to a form with terms proportional to (q), and .
Next, obtain (B) using (??) and (??). Finally, work out the last form
of (3.25) using the vector identity (??) to rearrange terms and =
B = (B ) /.] ///*

3.32 Show that the large aspect ratio tokamak expansion of the general expression
for local shear yields the result indicated after the ' in (3.119). //*

DRAFT 22:52
September 22, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

Chapter 5

Plasma Descriptions I:
Kinetic, Two-Fluid
Descriptions of plasmas are obtained from extensions of the kinetic theory of
gases and the hydrodynamics of neutral fluids (see Sections A.4 and A.6). They
are much more complex than descriptions of charge-neutral fluids because of
the complicating effects of electric and magnetic fields on the motion of charged
particles in the plasma, and because the electric and magnetic fields in the
plasma must be calculated self-consistently with the plasma responses to them.
Additionally, magnetized plasmas respond very anisotropically to perturbations
because charged particles in them flow almost freely along magnetic field
lines, gyrate about the magnetic field, and drift slowly perpendicular to the
magnetic field.
The electric and magnetic fields in a plasma are governed by the Maxwell
equations (see Section A.2). Most calculations in plasma physics assume that
the constituent charged particles are moving in a vacuum; thus, the microscopic, free space Maxwell equations given in (??) are appropriate. For some
applications the electric and magnetic susceptibilities (and hence dielectric and
magnetization responses) of plasmas are derived (see for example Sections 1.3,
1.4 and 1.6); then, the macroscopic Maxwell equations are used. Plasma effects
enter the Maxwell equations through the charge density and current sources
produced by the response of a plasma to electric and magnetic fields:
!
!
ns q s , J =
ns qs Vs , plasma charge, current densities. (5.1)
q =
s

Here, the subscript s indicates the charged particle species (s = e, i for electrons,
ions), ns is the density (#/m3 ) of species s, qs the charge (Coulombs) on the
species s particles, and Vs the species flow velocity (m/s). For situations where
the currents in the plasma are small (e.g., for low plasma pressure) and the
magnetic field, if present, is static, an electrostatic model (E = , E =
q /#0 = 2 = q /#0 ) is often appropriate; then, only the charge density
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

q is needed. The role of a plasma description is to provide a procedure for


calculating the charge density q and current density J for given electric and
magnetic fields E, B.
Thermodynamic or statistical mehanics descriptions (see Sections A.3 and
A.5) of plasmas are possible for some applications where plasmas are close to
a Coulomb collisional equilibrium. However, in general such descriptions are
not possible for plasmas because plasmas are usually far from a thermodynamic or statistical mechanics equilibrium, and because we are often interested
in short-time-scale plasma responses before Coulomb collisional relaxation processes become operative (on the 1/ time scale for fluid properties). Also, since
the lowest order velocity distribution of particles is not necessarily an equilibrium Maxwellian distribution, we frequently need a kinetic decsription to
determine the velocity as well as the spatial distribution of charged particles in
a plasma.
The pedagogical approach we employ in this Chapter begins from a rigorous microscopic description based on the sum of the motions of all the charged
particles in a plasma and then takes successive averages to obtain kinetic, fluid
moment and (in the next chapter) magnetohydrodynamic (MHD) descriptions
of plasmas. The first section, 5.1, averages the microscopic equation to develop
a plasma kinetic equation. This fundamental plasma equation and its properties
are explored in Section 5.2. [While, as indicated in (5.1), only the densities and
flows are needed for the charge and current sources in the Maxwell equations,
often we need to solve the appropriate kinetic equation and then take velocityspace averages of it to obtain the needed density and flow velocity of a particle
species.] Then, we take averages over velocity space and use various approximations to develop macroscopic, fluid moment descriptions for each species of
charged particles within a plasma (Sections 5.3*, 5.4*). The properties of a
two-fluid (electrons, ions) description of a magnetized plasma [e.g., adiabatic,
fluid (inertial) responses, and electrical resistivity and diffusion] are developed
in the next section, 5.5. Then in Section 5.6*, we discuss the flow responses in
a magnetized two-fluid plasma parallel, cross (EB and diamagnetic) and
perpendicular (transport) to the magnetic field. Finally, Section 5.7 discusses
the relevant time and length scales on which the kinetic and two-fluid models
of plasmas are applicable, and hence useful for describing various unmagnetized
plasma phenomena. This chapter thus presents the procedures and approximations used to progress from a rigorous (but extremely complicated) microscopic
plasma description to succesively more approximate (but progressively easier to
use) kinetic, two-fluid and MHD macroscopic (in the next chapter) descriptions,
and discusses the key properties of each of these types of plasma models.

5.1

Plasma Kinetics

The word kinetic means of or relating to motion. Thus, a kinetic description


includes the effects of motion of charged particles in a plasma. We will begin
from an exact (albeit enormously complicated), microscopic kinetic description
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

that is based on and encompasses the motions of all the individual charged particles in the plasma. Then, since we are usually interested in average rather than
individual particle properties in plasmas, we will take an appropriate average
to obtain a general plasma kinetic equation. Here, we only indicate an outline of the derivation of the plasma kinetic equation and some of its important
properties; more complete, formal derivations and discussions are presented in
Chapter 13.
The microsopic description of a plasma will be developed by adding up the
behavior and effects of all the individual particles in a plasma. We can consider
charged particles in a plasma to be point particles because quantum mechanical effects are mostly negligible in plasmas. Hence, the spatial distribution of
a single particle moving along a trajectory x(t) can be represented by the delta
function [x x(t)] = [x x(t)] [y y(t)] [z z(t)] see B.2 for a discussion of the spikey (Dirac) delta functions and their properties. Similarly, the
particles velocity space distribution while moving along the trajectory v(t) is
[vv(t)]. Here, x, v are Eulerian (fixed) coordinates of a six-dimensional phase
space (x, y, z, vx , vy , vz ), whereas x(t), v(t) are the Lagrangian coordinates that
move with the particle.
Adding up the products of these spatial and velocity-space delta function
distributions for each of the i = 1 to N (typically 1016 1024 ) charged particles
of a given species in a plasma yields the spikey microscopic (superscript m)
distribution for that species of particles in a plasma:
f (x, v, t) =
m

N
!
i=1

[x xi (t)] [v vi (t)],

microscopic distribution function.

(5.2)
The units of a distribution function are the reciprocal of the volume in the sixdimensional phase space x, v or # /(m6 s3 ) recall that the units of a delta
function are one over the units of its argument (see B.2). Thus, d3x d3v f is
the number of particles in the six-dimensional phase space differential volume
between x, v and x+dx, v +dv. The distribution function in (5.2) is normalized
such that its integral over velocity space yields the particle density:
n (x, t)
m

"

d v f (x, v, t) =
3

N
!
i=1

[x xi (t)],

particle density (#/m3 ).

(5.3)
Like the distribution f m , this microscopic density distribution is very singular
or spikey it is infinite at the instantaneous particle positions x = xi (t) and
zero elsewhere. Integrating the density over the volume V of
# the plasma yields
the total number of this species of particles in the plasma: V d3x n(x, t) = N .
Particle trajectories xi (t), vi (t) for each of the particles are obtained from
their equations of motion in the microscopic electric and magnetic fields Em , Bm :
m dvi /dt = q [Em (xi , t) + vi Bm (xi , t)],

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

dxi /dt = vi ,

i = 1, 2, . . . , N.
(5.4)

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

(The portion of the Em , Bm fields produced by the ith particle is of course omitted from the force on the ith particle.) In Eqs. (5.2)(5.4), we have suppressed
the species index s (s = e, i for electrons, ions) on the distribution function f m ,
the particle mass m and the particle charge q; it will be reinserted when needed,
particularly when summing over species.
The microscopic electric and magnetic fields Em , Bm are obtained from the
free space Maxwell equations:
m
m
Em = m
q /#0 , E = B /t,

Bm = 0,

Bm = 0 (Jm + #0 Em /t).

(5.5)

The required microscopic charge and current sources are obtained by integrating
the distribution function over velocity space and summing over species:
m
q (x, t)
J (x, t)
m

qs

!
s

qs

"

"

d3v fsm (x, v, t) =

qs

v vfsm (x, v, t)

i=1

!
s

N
!

qs

[x xi (t)],

N
!
i=1

(5.6)

vi (t) [x xi (t)].

Equations (5.2)(5.6) together with initial conditions for all the N particles
provide a complete and exact microscopic description of a plasma. That is,
they describe the exact motion of all the charged particles in a plasma, their
consequent charge and current densities, the electric and magnetic fields they
generate, and the effects of these microscopic fields on the particle motion all
of which must be calculated simultaneously and self-consistently. In principle,
one can just integrate the N particle equations of motion (5.4) over time and
obtain a complete description of the evolving plasma. However, since typical
plasmas have 1016 1024 particles, this procedure involves far too many equations to ever be carried out in practice1 see Problem 5.1. Also, since this
description yields the detailed motion of all the particles in the plasma, it yields
far more detailed information than we need for practical purposes (or could
cope with). Thus, we need to develop an averaging scheme to reduce this microscopic description to a tractable set of equations whose solutions we can use
to obtain physically measurable, average properties (e.g., density, temperature)
of a plasma.
To develop an averaging procedure, it would be convenient to have a single
evolution equation for the entire microscopic distribution f m rather than having
1 However, particle-pushing computer codes carry out this procedure for up to millions
of scaled macro particles. The challenge for such codes is to have enough particles in each
relevant phase space coordinate so that the noise level in the simulation is small enough
to not mask the essential physics of the process being studied. High fidelity simulations
are often possible for reduced dimensionality applications. Some relevant references for this
fundamental computational approach are: J.M. Dawson, Rev. Mod. Phys. 55, 403 (1983);
C.K. Birdsall and A.B. Langdon, Plasma Physics Via Computer Simulation (McGraw-Hill,
New York, 1985); R.W. Hockney and J.W. Eastwood, Computer Simulation Using Particles
(IOP Publishing, Bristol, 1988).

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

to deal with a very large number (N ) of particle equations of motion. Such an


equation can be obtained by calculating the total time derivative of (5.2):
df m
dt

% N

dx
dv !

[x xi (t)] [v vi (t)]
t
dt x
dt v i=1
%
N $
!
dxi
dvi

[x xi (t)] [v vi (t)]
=
t
dt x
dt v
i=1
N $
!
dvi
dxi

=
dt x
dt v
i=1
%
dvi
dxi

[x xi (t)] [v vi (t)]
+
dt x
dt v
= 0.
(5.7)

Here in successive lines we have used three-dimensional forms of the properties


of delta functions given in (??), and (??): x (x xi ) = xi (x xi ) and v (v
vi ) = vi (v vi ), and (/t) [x xi (t)] = dxi /dt (/x) [x xi (t)] and
(/t) [v vi (t)] = dvi /dt (/v) [v vi (t)]. Substituting the equations of
motion given in (5.4) into the second line of (5.7) and using the delta functions
to change the functional dependences of the partial derivatives from xi , vi to
x, v, we find that the result df m /dt = 0 can be written in the equivalent forms
df m
dt

f m
dx f m
dv f m
+

t
dt x
dt v
m
m
f
q
f m
f
+v
+ [Em (x, t) + vBm (x, t)]
= 0. (5.8)
t
x
m
v

This is called the Klimontovich equation.2 Mathematically, it incorporates all N of the particle equations of motion into one equation because the
mathematical characteristics of this first order partial differential equation in
the seven independent, continuous variables x, v, t are dx/dt = v, dv/dt =
(q/m)[Em (x, t) + vBm (x, t)], which reduce to (5.4) at the particle positions:
x xi , v vi for i = 1, 2, . . . , N . That is, the first order partial differential equation (5.8) advances positions in the six-dimensional phase space x, v
along trajectories (mathematical characteristics) governed by the single particle
equations of motion, independent of whether there is a particle at the particular
phase point x, v; if say the ith particle is at this point (i.e., x = xi , v = vi ),
then the trajectory (mathematical characteristic) is that of the ith particle.
Equations (5.2), (5.5), (5.6) and (5.8) provide a complete, exact description
of our microscopic plasma system that is entirely equivalent to the one given
by (5.2)(5.6); this Klimintovich form of the equations is what we will average
below to obtain our kinetic plasma description. These and other properties of
the Klimontovich equation are discussed in greater detail in Chapter 13.
2 Yu. L. Klimontovich, The Statistical Theory of Non-equilibrium Processes in a Plasma
(M.I.T. Press, Cambridge, MA, 1967); T.H. Dupree, Phys. Fluids 6, 1714 (1963).

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

The usual formal procedure for averaging a microscopic equation is to take


its ensemble average.3 We will use a simpler, more physical procedure. We begin
by defining the number of particles N6D in a small box in the six-dimensional
(6D) phase space of spatial volume
# V # x y z and velocity-space volume
Vv vx vy vz : N6D V d3x Vv d3v f m . We need to consider box
sizes that are large compared to the mean spacing of particles in the plasma [i.e.,
x >> n1/3 in physical space and vx >> vT /(n3D )1/3 in velocity space] so
there are many particles in the box and hence the statistical fluctuations
in the

number of particles in the box will be small (N6D /N6D 1/ N6D << 1).
However, it should not be so large that macroscopic properties of the plasma
(e.g., the average density) vary significantly within the box. For plasma applications the box size should generally be smaller than, or of order the Debye length
D for which N6D (n3D )2 >>>> 1 so collective plasma responses on the
Debye length scale can be included in the analysis. Thus, the box size should be
large compared to the average interparticle spacing but small compared to the
Debye length, a criterion which will be indicated in its one-dimensional spatial
form by n1/3 < x < D . Since n3D >> 1 is required for the plasma state, a
large range of xs fit within this inequality range.
The average distribution function )f m * will be defined as the number of
particles in such a small six-dimensional phase space box divided by the volume
of the box:
#
#
d3x Vv d3v f m
N6D
V
m
#
#
lim
=
lim
,
)f (x, v, t)*
d3x Vv d3v
n1/3 <x<D V Vv
n1/3 <x<D
V
average distribution function. (5.9)

From this form it is clear that the units of the average distribution function are
the number of particles per unit volume in the six-dimensional phase space, i.e.,
#/(m6 s3 ). In the next section we will identify the average distribution )f m *
as the fundamental plasma distribution function f .
The deviation of the complete microscopic distribution f m from its average,
which by definition must have zero average, will be written as f m :
discrete particle distribution function.
(5.10)
The average distribution function )f m * represnts the smoothed properties of the
> D ; the microscopic distribution f m represents the
plasma species for x
< x < D .
discrete particle effects of individual charged particles for n1/3
This averaging procedure is illustrated graphically for a one-dimensional
system in Fig. 5.1. As indicated, the microscopic distribution f m is spikey
because it represents the point particles by delta functions. The average
distribution function )f m * indicates the average number of particles over length
f m f m )f m *,

)f m * = 0,

3 In an ensemble average one obtains expectation values by averaging over an infinite


number of similar plasmas (realizations) that have the same number of particles and macroscopic parameters (e.g., density n, temperature T ) but whose particle positions vary randomly
(in the six-dimensional phase space) from one realization to the next.

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

Figure 5.1: One-dimensional illustration of the microscopic distribution function


f m , its average )f m * and its particle discreteness component f m .
scales that are large compared to the mean interparticle spacing. Finally, the
discrete particle distribution function f m is spikey as well, but has a baseline
of )f m (x)*, so that its average vanishes.
In addition to splitting the distribution function into its smoothed and discrete particle contributions, we need to split the electric and magnetic fields,
and charge and current densities into their smoothed and discrete particle parts
components:
Em
m
q

= )Em * + Em , Bm
m
= )m
q * + q ,

Jm

= )Bm * + Bm ,
= )Jm * + Jm .

(5.11)

Substituting these forms into the Klimontovich equation (5.8) and averaging
the resultant equation using the averaging definition in (5.9), we obtain our
fundamental plasma kinetic equation:
)f m *
q
)f m *
)f m *
+v
+ [)Em * + v)Bm *]
=
t
x
m
v
&
'
f m
q
[Em + vBm ]
. (5.12)

m
v
The terms on the left describe the evolution of the smoothed, average distribution function in response to the smoothed, average electric and magnetic fields
in the plasma. The term on the right represents the two-particle correlations
between discrete charged particles within about a Debye length of each other. In
fact, as can be anticipated from physical considerations and as will be shown in
detail in Chapter 13, the term on the right represents the small Coulomb collision effects on the average distribution function )f m *, whose basic effects were
calculated in Chapter 2. Similarly averaging the microscopic Maxwell equations (5.5) and charge and current density sources in (5.6), we obtain smoothed,
average equations that have no extra correlation terms like the right side of
(5.12).

5.2

Plasma Kinetic Equations

We now identify the smoothed, average [defined in (5.9)] of the microscopic


distribution function )f m * as the fundamental distribution function f (x, v, t)
for a species of charged particles in a plasma. Similarly, the smoothed, average
of the microscopic electric and magnetic fields, and charge and current densities
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

will be written in their usual unadorned forms: )Em * E, )Bm * B, )m


q *
q , and )Jm * J. Also, we write the right side of (5.12) as C(f ) a Coulomb
collision operator on the average distribution function f which will be derived
and discussed in Chapter 11. With these specifications, (5.12) can be written
as
f
f
q
f
df
=
+v
+ [E + vB]
= C(f ), f = f (x, v, t),
dt
t
x m
v
PLASMA KINETIC EQUATION. (5.13)
This is the fundamental plasma kinetic equation4 we will use thoughout the
remainder of this book to provide a kinetic description of a plasma. To complete
the kinetic description of a plasma, we also need the average Maxwell equations,
and charge and current densities:
(
)
B
E
q
, B = 0, B = 0 J + #0
. (5.14)
E = , E =
#0
t
t
! "
! "
qs d3v fs (x, v, t), J(x, t)
qs d3v vfs (x, v, t). (5.15)
q (x, t)
s

Equations (5.13)(5.15) are the fundamental set of equations that provide a


complete kinetic description of a plasma. Note that all of the quantities in
them are smoothed, average quantities that have been averaged according to
the prescription in (5.9). The particle discreteness effects (correlations of particles due to their Coulomb interactions within a Debye sphere) in a plasma are
manifested in the Coulomb collsion operator on the right of the plasma kinetic
equation (5.13). In the averaging procedure we implicitly assume that the particle discreteness effects do not extend to distances beyond the Debye length
D . Chapter 13 discusses two cases (two-dimensional magnetized plasmas and
convectively unstable plasmas) where this assumption breaks down. Thus, while
we will hereafter use the average plasma kinetic equation (5.13) as our fundamental kinetic equation, we should keep in mind that there can be cases where
the particle discreteness effects in a plasma are not completely represented by
the Coulomb collision operator.
For low pressure plasmas where the plasma currents are negligible and the
magnetic field (if present) is constant in time, we can use an electrostatic approximation for the electric field (E = ). Then, (5.13)(5.15) reduce to
f
q
f
f
+v
+ [ + vB]
= C(f ),
t
x m
v

(5.16)

4 Many plasma physics books and articles refer to this equation as the Boltzmann equation, thereby implicity indicating that the appropriate collision operator is the Boltzmann
collision operator in (??). However, the Coulomb collision operator is a special case (small
momentum transfer limit see Chapter 11) of the Boltzmann collision operator CB , and
importantly involves the cumulative effects (the ln factor) of multiple small-angle, elastic
Coulomb collsions within a Debye sphere that lead to diffusion in velocity-space. Also, the
Boltzmann equation usually does not include the electric and magnetic field effects on the
charged particle trajectories during collisions or on the evolution of the distribution function.
Thus, this author thinks it is not appropriate to call this the Boltzmann equation.

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID


q
= ,
#0
2

q =

qs

"

d3vf (x, v, t),

9
(5.17)

which provides a complete electrostatic, kinetic description of a plasma.


Some alternate forms of the general plasma kinetic equation (5.13) are also
useful. First, we derive a conservative form of it. Since x and v are independent Eulerian phase space coordinates, using the vector identity (??) we
find
(
)
f

vf = v
+f
v = v
.
x
x
x
x

Similarly, for the velocity derivative we have

q
q
f

[E + vB] f = [E + vB]
,
v m
m
v
since /v [E + vB] = 0 because E, B are both independent of v, and
/v vB = 0 using vector identities (??) and (??). Using these two results
we can write the plasma kinetic equation as
+

*q
f
+
[vf ] +

(E + vB) f
= C(f ),
t x
v
m
conservative form of plasma kinetic equation,(5.18)
which is similar to the corresponding neutral gas kinetic equation (??). Like for
the kinetic theory of gases, we can put the left side of the plasma kinetic equation
in a conservative form because (in the absence of collisions) motion (of particles
or along the characteristics) is incompressible in the six-dimensional phase space
x, v: /x (dx/dt) + /v (dv/dt) = /x v + /v (q/m)[E + vB] = 0
see (??).
In a magnetized plasma with small gyroradii compared to perpendicular
gradient scale lengths (( << 1) and slow processes compared to the gyrofrequency (/t << c ), it is convenient to change the independent phase space
variables from x, v phase space to the guiding center coordinates xg , g , . (The
third velocity-space variable would be the gyromotion angle , but that is averaged over to obtain the guiding center motion equations see Section 4.4.)
Recalling the role of the particle equations of motion (5.4) in obtaining the
Klimintovich equation, we see that in terms of the guiding center coordinates
the plasma kinetic equation becomes f /t + dxg /dt f + (d/dt) f / +
(dg /dt) f / g = C(f ). The gyroaverage of the time derivative of the magnetic moment and f / are both small in the small gyroradius expansion;
hence their product can be neglected in this otherwise first order (in a small
gyroradius expansion) plasma kinetic equation. The time derivative of the energy can be calculated to lowest order (neglecting the drift velocity vD ) using
the guiding center equation (??), writing the electric field in its general form
E = A/t and d/dt = /t + dxg /dt /x + /t + v# # :
,
2
d mv#
d
dB

B
d g
A .
=
+q
+
+ q
+
qv# b
(5.19)
dt
dt
2
dt
dt
t
t
t
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

10

Thus, after averaging the plasma kinetic equation over the gyromotion angle ,
the plasma kinetic equation for the gyro-averaged, guiding-center distribution
function fg can be written in terms of the guiding center coordinates (to lowest
order neglecting vD# ) as
fg
fg + vD fg + dg fg
+ v# b
t
dt g

= )C(fg )* ,

fg = f (xg , g , , t),

drift-kinetic equation, (5.20)


in which the collision operator is averaged over gyrophase [see discussion before
(??)] and the spatial gradient is taken at constant g , , t, i.e., /x |g ,,t .
This lowest order drift-kinetic equation is sufficient for most applications. However, like the guiding center orbits it is based on, it is incorrect at second order
in the small gyroradius expansion [for example, it cannot be put in the conservative form of (5.18) or (??)]. More general and accurate gyrokinetic equations
that include finite gyroradius effects (( 1) have also been derived; they
are used when more precise and complete equations are needed.
For many plasma processes we will be interested in short time scales during
which Coulomb collision effects are negligible. For these situations the plasma
kinetic equation becomes
f
f
q
f
df
=
+v
+ [E + vB]
= 0,
dt
t
x m
v
Vlasov equation.

(5.21)

This equation, which is also called the collisionless plasma kinetic equation,
was originally derived by Vlasov5 by neglecting the particle discreteness effects
that give rise to the Coulomb collisional effects see Problem 5.2. Because
the Vlasov equation has no discrete particle correlation (Coulomb collision)
effects in it, it is completely reversible (in time) and its solutions follow the
collisionless single particle orbits in the six-dimensional phase space. Thus, its
distribution function solutions are entropy conserving (there is no irreversible
relaxation of irregularities in the distribution function), and, like the particle
orbits, incompresssible in the six-dimensional phase space see Section 13.1.
The nominal condition for the neglect of collisional effects is that the frequency of the relevant physical process(es) be much larger than the collision
frequency: d/dt i >> , in which is the Lorentz collision frequency
(??). Here, the frequency represents whichever of the various fundamental frequencies (e.g., p , plasma; kcS , ion acoustic; c , gyrofrequency; b ,
bounce; D , drift) are relevant for a particular plasma application. However,
since the Coulomb collision process is diffusive in velocity space (see Section
2.1 and Chapter 11), for processes localized to a small region of velocity space
v /v << 1, the effective collision frequency (for scattering out of this
narrow region of velocity space) is eff /2 >> . For this situation the
relevant condition for validity of the Vlasov equation becomes >> eff . Often,
5 A.A.

Vlasov, J. Phys. (U.S.S.R.) 9, 25 (1945).

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

11

the Vlasov equation applies over most of velocity space, but collisions must be
taken into account to resolve singular regions where velocity-space derivatives
of the collisionless distribution function are large.
Finally, we briefly consider equilibrium solutions of the plasma kinetic and
Vlasov equations. When the collision operator is dominant in the plasma kinetic equation (i.e., >> ), the lowest order distribution is the Maxwellian
distribution [see Chapter 11 and (??)]:
fM (x, v, t) = n

(
)
2
2
. m /3/2
nevr /vT
m|vr |2
= 3/2 3 , vr v V,
exp
2T
2T
vT
Maxwellian distribution function. (5.22)

0
Here, vT 2T /m is the thermal velocity, which is the most probable speed
[see (??)] in the Maxwellian distribution. Also, n(x, t) is the density (units of
#/m3 ), T (x, t) is the temperature (J or eV) and V(x, t) the macroscopic flow
velocity (m/s) of the species of charged particles being considered. Note that
the vr in (5.22) represents the velocity of a particular particle in the Maxwellian
distribution relative to the #average macroscopic flow velocity of the entire distribution of particles: V d3v vfM /n. It can be shown (see Chapter 13) that
the collisionally relaxed Maxwellian distribution has no free energy in velocity
space to drive (kinetic) instabilities (collective fluctuations whose magnitude
grows monotonically in time) in a plasma; however, its spatial gradients (e.g.,
n and T ) provide spatial free energy sources that can drive fluidlike (as
opposed to kinetic) instabilities see Chapters 2123.
If collisions are negligible for the processes being considered (i.e., >> eff ),
the Vlasov equation is applicable. When there exist constants of the single
particle motion ci (e.g., energy c1 = , magnetic moment c2 = , etc. which
satisfy dci /dt = 0), solutions of the Vlasov equation can be written in terms of
them:
f

f (c1 , c2 , ), ci = constants of motion,


! dci f
df
=
= 0.
=
dt
dt ci
i
=

Vlasov equation solution,


(5.23)

A particular Vlasov solution of interest is when the energy is a constant


of the motion and the equilibrium distribution function depends only on it:
f0 = f0 (). If such a distribution is a monotonically decreasing function of the
energy (i.e., df0 /d < 0), then one can readily see from physical considerations
and show mathematically (see Section 13.1) that this equilibrium distribution
function has no free energy available to drive instabilities because all possible
rearrangements of the energy distribution, which must be area-preserving in the
six-dimensional phase space
because
of the Vlasov equation df /dt = 0, would
#
#
raise the system energy d3x d3v (mv 2 /2)f () leaving no free energy available
to excite unstable electric or magnetic fluctuations. Thus, we have the statement
f0 = f0 (), with df0 /d < 0, is a kinetically stable distribution.
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

(5.24)

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

12

Note that the Maxwellian distribution in (5.22) satisfies these conditions if there
are no spatial gradients in the plasma density, temperature or flow velocity.
However, confined plasmas must have additional dependencies on spatial coordinates6 or constants of the motion so they can be concentrated in regions
within and away from the plasma boundaries. Thus, most plasmas of interest do not satisfy (5.24). The stability of such plasmas has to be investigated
mostly on a case-by-case basis. When instabilities occur they usually provide
the dominant mechanisms for relaxing plasmas toward a stable (but unconfined
plasma) distribution function of the type given in (5.24).

5.3

Fluid Moments*

For many plasma applications, fluid moment (density, flow velocity, temperature) descriptions of a charged particle species in a plasma are sufficient. This is
generally the case when there are no particular velocities or regions of velocity
space where the charged particles behave differently from the typical thermal
particles of that species. In this section we derive fluid moment evolution equations by calculating the physically most important velocity-space moments of the
plasma kinetic equation (density, momentum and energy) and discuss the closure moments needed to close the fluid moment hierarchy of equations. This
section is mathematically intensive with many physical details for the various
fluid moments; it can skipped since the key features of fluid moment equations
for electrons and ions are summarized at the beginning of the section after the
next one.
Before beginning the derivation of the fluid moment equations, it is convenient to define the various velocity moments of the distribution function we
will need. The various moments result from integrating low order powers of the
velocity#v times the distribution function f over velocity space in the laboratory
frame: d3v vj f, j = 0, 1, 2. The integrals are all finite because the distribution
function must fall off sufficiently rapidly with speed so that these low order,
physical moments (such as the energy in the species) are finite. That is, we
cannot have large numbers of particles at arbitrarily high energy because then
the energy in the species would be unrealistically large or divergent. [Note that
velocity integrals of all algebraic powers of the velocity times the Maxwellian
distribution (5.22) converge see Section C.2.] The velocity moments of the
distribution function f (x, v, t) of physical interest are
"
n d3v f,
(5.25)
density (#/m3 ) :
"
1
d3v vf,
(5.26)
flow velocity (m/s) :
V
n
6 One could use the potential energy term q(x) in the energy to confine a particular species
of plasma particles but the oppositely charged species would be repelled from the confining
region and thus the plasma would not be quasineutral. However, nonneutral plasmas can be
confined by a potential .

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID


temperature (J, eV) :
conductive heat flux (W/m2 ) :
pressure (N/m2 ) :
pressure tensor (N/m2 ) :
stress tensor (N/m2 ) :

1
n
"

"

mvT2
mvr2
f=
,
3
2
)
(
mvr2
f,
q d3v vr
2
"
mvr2
f = nT,
p d3v
3
"
P d3v mvr vr f = pI + ,
(
)
"
v2
d3v m vr vr r I f,
3
T

d3v

13
(5.27)
(5.28)
(5.29)
(5.30)
(5.31)

in which we have defined and used


vr |vr |, relative (subscript r) velocity, speed. (5.32)
#
By definition, we have d3v vr f = n(V V) = 0. For simplicity, the species
subscript s = e, i is omitted here and thoughout most of this section; it is
inserted only when needed to clarify differences in properties of electron and ion
fluid moments.
All these fluid moment properties of a particular species s of charged particles
in a plasma are in general functions of spatial position x and time t: n = n(x, t),
etc. The density n is just the smoothed average of the microscopic density (5.3).
The flow velocity V is the macroscopic flow velocity of this species of particles.
The temperature T is the average energy of this species of particles, and is
measured in the rest frame of this species of particles hence the integrand
is (mvr2 /2)f instead of (mv 2 /2)f . The conductive heat flux q is the flow of
energy density, again measured in the rest frame of this species of particles. The
pressure p is a scalar function that represents the isotropic part of the expansive
stress (pI in P in which I is the identity tensor) of particles since their thermal
motion causes them to expand isotropically (in the species rest frame) away
from their initial positions. This is an isotropic expansive stress on the species
of particles because the effect of the thermal motion of particles in an isotropic
distribution is to expand uniformly in all directions; the net force (see below)
due to this isotropic expansive stress is pI = I pp I = p (in the
direction from high to low pressure regions), in which the vector, tensor identities
(??), (??) and (??) have been used. The pressure tensor P represents the overall
pressure stress in the species, which can have both isotropic and anisotropic
(e.g., due to flows or magnetic field effects) stress components. Finally, the
stress tensor is a traceless, six-component symmetric tensor that represents
the anisotropic components of the pressure tensor.
In addition, we will need the lowest order velocity moments of the Coulomb
collision operator C(f ). The lowest order forms of the needed moments can be
inferred from our discussion of Coulomb collisions in Section 2.3:
"
(5.33)
density conservation in collisions :
0 = d3v C(f ),
vr v V(x, t),

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID


frictional force density (N/m ) :
3

energy exchange density (W/m3 ) :

R
Q

"

"

d3v mv C(f ),
d3v

mvr2
C(f ).
2

14
(5.34)
(5.35)

As indicated in the first of these moments, since Coulomb collisions do not


create or destroy charged particles, the density moment of the collision operator vanishes. The momentum moment of the Coulomb collision operator
represents the (collisional friction) momentum gain or loss per unit volume
from a species of charged particles that is flowing relative to another species:
Re + me ne e (Ve Vi ) = ne eJ/ and Ri = Re from (??) and (??). Here,
rigorously speaking, the electrical conductivity is the Spitzer value (??). (The
approximate equality here means that we are neglecting the typically small effects due to temperature gradients that are needed for a complete, precise theory
see Section 12.2.) The energy moment of the collision operator represents the
rate of Coulomb collisional energy exchange per unit volume between two species
of charged particles of different temperatures: Qi = 3(me /mi )e ne (Te Ti ) and
Qe + J 2 / Qi from (??) and (??). In a magnetized plasma, the electrical
conductivity along the magnetic field is the Spitzer value [# = Sp from (??)],
but perpendicular to the magnetic field it is the reference conductivity [ = 0
from (??)] (because the gyromotion induced by the B field impedes the perpendicular motion and hence prevents the distortion of the distribution away from
a flow-shifted Maxwellian see discussion near the end of Section 2.2 and in
# /# + J / ) and
Section 12.2). Thus, in a magnetized plasma Re = ne e(bJ
2
/ Qi .
Qe = J#2 /# + J
As in the kinetic theory of gases, fluid moment equations are derived by
taking velocity-space moments of a relevant kinetic equation, for which it is
simplest to use the conservative form (5.18) of the plasma kinetic equation:
$
%
"

q
f
+
vf +
(E + vB)f C(f ) = 0
(5.36)
d3v g(v)
t x
v m
in which g(v) is the relevant velocity function for the desired fluid moment.
We begin by obtaining the density moment by evaluating (5.36) using g = 1.
Since the Eulerian velocity space coordinate v is stationary and hence is independent of time, the time derivative can be interchanged with the integral
over
#
velocity space. (Mathematically, the partial time derivative and d3v operators commute, i.e.,
# their order can be interchanged.) # Thus, the first integral
becomes (/t) d3v f = n/t. Similarly, since the d3v and spatial derivative /x operators commute, they# can be interchanged in the second term in
(5.36) which then becomes /x d3v vf = /x nV nV. Since the
integrand in the third term in (5.36) is in the form of a divergence in velocity
space, its #integral can be converted
# into a surface integral using Gauss theorem (??): d3v /v (dv/dt)f = dSv (dv/dt)f = 0, which vanishes because
there must be exponentially few particles on the bounding velocity space surface |v| so that all algebraic moments of the distribution function are
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

15

finite and hence exist. Finally, as indicated in (5.33) the density moment of the
Coulomb collision operator vanishes.
Thus, the density moment of the plasma kinetic equation yields the density
continuity or what is called simply the density equation:
dn
= n V.
dt
(5.37)
Here, in obtaining the second form we used the vector identity (??) and the last
form is written in terms of the total time derivative (local partial time derivative
plus that induced by advection7 see Fig. 5.2a below) in a fluid moving with
flow velocity V:
1
11
d

+ V , total time derivative in a moving fluid.


(5.38)
dt
t 1x
n
+ nV = 0
t

n
= V n n V
t

This total time derivative is sometimes called the substantive derivative. From
the middle form of the density equation (5.37) we see that at a fixed (Eulerian)
position, increases (n/t > 0) in the density of a plasma species are caused
by advection of the species at flow velocity V across a density gradient from
a region of higher density into the local one with lower density (V n <
0), or by compression ( V < 0, convergence) of the flow. Conversely, the
local density decreases if the plasma species flows from a lower into a locally
higher density region or if the flow is expanding (diverging). The last form
in (5.37) shows that in a frame of reference moving with the flow velocity V
(Lagrangian description) only compression (expansion) of the flow causes the
density to increase (decrease) see Fig. 5.2b below.
The momentum equation for a plasma species is derived similarly by taking
the momentum moment of the plasma kinetic equation. Using g = mv in
(5.36), calculating the various terms as in the preceding paragraph and using
vv = (vr + V)(vr + V) in evaluating the second term, we find
m (nV)/t + (pI + + mnVV) nq [E + VB] R = 0.

(5.39)

In obtaining the next to last term we have used vector identity (??) to write
v /v [(dv/dt)f ] = /v [v(dv/dt)f ] (dv/dt)f (v/v), which is equal
to /v [v(dv/dt)f ] (dv/dt)f since v/v I and dv/dt I = dv/dt; the
term containing the divergence in velocity space again vanishes by conversion
to a surface integral, in this case using (??). Next we rewrite (5.38) using (5.37)
to remove the n/t contribution and mnVV = mV( nV) + mnV V
to obtain
dV
= nq [E + VB] p + R
(5.40)
mn
dt
in which the total time derivative d/dt in the moving fluid is that defined in
(5.38). Equation (5.40) represents the average of Newtons second law (ma = F)
7 Many plasma physics books and articles call this convection. In fluid mechanics advection
means transport of any quantity by the flow velocity V and convection refers only to the heat
flow (5/2)nT V induced by the fluid flow. This book adopts the terminology of fluid mechanics.

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

16

over an entire distribution of particles. Thus, the mn dV/dt term on the left
represents the inertial force per unit volume in this moving (with flow velocity
V) charged particle species. The first two terms on the right give the average
(over the distribution function) force density on the species that results from
the Lorentz force q [E + vB] on the charged particles. The next two terms
represent the force per unit volume on the species that results from the pressure
tensor P = pI + , i.e,, both that due to the isotropic expansive pressure p and
the anisotropic stress . The R term represents the frictional force density on
this species that results from Coulomb collisional relaxation of its flow V toward
the flow velocities of other species of charged particles in the plasma.
Finally, we obtain the energy equation for a plasma species by taking the
energy moment of the plasma kinetic equation. Using g = mv 2 /2 in (5.36) and
proceeding as we did for the momentum moment, we obtain (see Problem 5.??)
(
)
$
(
)
%
1
1
5
3
2
2
nT + mnV
nT + nmV
+ q +
V+V
t 2
2
2
2
nqV E Q V R = 0.
(5.41)
Using the dot product of the momentum equation (5.40) with V to remove
the V 2 /t term in this equation and using the density equation (5.37), this
equation can be simplified to
(
)
5
3 p
= q + pV + V p : V + Q,
2 t
2
3 dp 5
+ p V = q : V + Q
(5.42)
or,
2 dt
2
The first form of the energy equation shows that the local (Eulerian) rate of increase of the internal energy per unit volume of the species [(3/2)nT = (3/2)p]
is given by the sum of the net (divergence of the) energy fluxes into the local
volume due to heat conduction (q), heat convection [(5/2)pV (3/2)pV internal energy carried along with the flow velocity V plus pV from mechanical
work done on or by the species as it moves], advection of the pressure from a
lower pressure region into the local one of higher pressure (V p > 0), and
dissipation due to flow-gradient-induced stress in the species ( : V) and
collisional energy exchange (Q).
The energy equation is often written in the form of an equation for the
time derivative of the temperature. This form is obtained by using the density
equation (5.37) to eliminate the n/t term implicit in p/t in (5.42) to yield
3 dT
n
= nT ( V) q : V + Q,
2 dt

(5.43)

in which d/dt is the total time derivative for the moving fluid defined in (5.38).
This form of the energy equation shows that the temperature T of a plasma
species increases (in a Lagrangian frame moving with the flow velocity V)
due to a compressive flow ( V < 0), the divergence of the conductive heat
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

17

flux ( q), and dissipation due to flow-gradient-induced stress in the species


( : V) and collisional energy exchange (Q).
Finally, it often useful to switch from writing the energy equation in terms
of the temperature or pressure to writing it in terms of the collisional entropy.
The (dimensionless) collisional entropy s for f + fM is
( 3/2 )
"
3 . p /
T
1
3
d v f ln f + ln
+ C = ln 5/3 + C, collisional entropy,
s
n
n
2
n
(5.44)
in which C is an unimportant constant. Entropy represents the state of disorder
of a system see the discussion at the end of Section A.3. Mathematically,
it is the logarithm of the number of number of statistically independent states
a particle can have in a relevant volume in the six-dimensional phase space.
For classical (i.e., non-quantum-mehanical) systems, it is the logarithm of the
average volume of the six-dimensional phase space occupied by one particle.
That is, it is the logarithm of the inverse of the density of particles in the
six-dimensional phase space, which for the collisional equilibrium Maxwellian
distribution (5.22) is + 3/2 vT3 /n T 3/2 /n. Entropy increases monotonically
in time as collisions cause particles to spread out into a larger volume (and
thereby reduce their density) in the six-dimensional phase space, away from an
originally higher density (smaller volume, more confined) state.
An entropy equation can be obtained directly by using the density and energy
equations (5.37) and (5.43) in the total time derivative of the entropy s for a
given species of particles:
nT

3 dT
dn
ds
= n
T
= q : V + Q.
dt
2 dt
dt

(5.45)

Increases in entropy (ds/dt > 0) in the moving fluid are caused by net heat
flux into the volume, and dissipation due to flow-gradient-induced stress in the
species and collisional energy exchange. The evolution of entropy in the moving
fluid can be written in terms of the local time derivative of the entropy density
ns by making use of the density equation (5.37) and vector identity (??):
$
%
$
%
d(ns)
dn
(ns)
ds
=T
s
=T
+ nsV .
(5.46)
nT
dt
dt
dt
t
Using this form for the rate of entropy increase and (q/T ) = (1/T )[ q
q ln T ] in (5.45), we find (5.45) can be written
.
q/
1
(ns)
+ nsV +
= (q ln T + : V Q).
(5.47)
t
T
T

In this form we see that local temporal changes in the entropy density [(ns)/t]
plus the net (divergence of) entropy flow out of the local volume by entropy convection (nsV) and heat conduction (q/T ) are induced by the dissipation in the
species (), which is caused by temperature-gradient-induced conductive heat
flow [q ln T = (1/T )q T ], flow-gradient-induced stress ( : V),
and collisional energy exchange (Q).
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

18

The fluid moment equations for a charged plasma species given in (5.37),
(5.40) and (5.43) are similar to the corresponding fluid moment equations obtained from the moments of the kinetic equation for a neutral gas (??)(??).
on a plasma species
The key differences are that: 1) the average force density nF
is given by the Lorentz force density n[E+VB] instead the gravitational force
mVG ; and 2) the effects of Coulomb collisions between different species of
charged particles in the plasma lead to frictional force (R) and energy exchange
(Q) additions to the momentum and energy equations. For plasmas there is of
course the additional complication that the densities and flows of the various
species of charged particles in a plasma have to be added according to (5.1)
to yield the charge q and current J density sources for the Maxwell equations
that then must be solved to obtain the E, B fields in the plasma, which then
determine the Lorentz force density on each species of particles in the plasma.
It is important to recognize that while each fluid moment of the kinetic
equation is an exact equation, the fluid moment equations represent a hierarchy
of equations which, without further specification, is not a complete (closed)
set of equations. Consider first the lowest order moment equation, the density
equation (5.37). In principle, we could solve it for the evolution of the density
n in time, if the species flow velocity V is specified. In turn, the flow velocity
is determined from the next order equation, the momentum equation (5.40).
However, to solve this equation for V we need to know the species pressure
(p = nT ) and hence really the temperature T , and the stress tensor . The
temperature is obtained from the isotropic version of the next higher order
moment equation, the energy equation (5.43). However, this equation depends
on the heat flux q.
Thus, the density, momentum and energy equations are not complete because we have not yet specified the highest order, closure moments in these
equations the heat flux q and the stress tensor . To determine them,
we could imagine taking yet higher order moments of the kinetic equation
[g = v(mvr2 /2) and m(vr vr (vr2 /3)I) in (5.36) ] to obtain evolution equations for q, . However, these new equations would involve yet higher order
moments (vvv, v 2 vv), most of which do not have simple physical interpretations and are not easily measured. Will this hierarchy never end?! Physically,
the even higher order moments depend on ever finer scale details of the distribution function f ; hence, we might hope that they are unimportant or negligible,
particularly taking account of the effects of Coulomb collisions in smoothing
out fine scale features of the distribution function in velocity-space. Also, since
the fluid moment equations we have derived so far provide evolution equations
for the physically most important (and measurable) properties (n, V, T ) of a
plasma species, we would like to somehow close the hierarchy of fluid moment
equations at this level.

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

5.4

19

Closure Moments*

The general procedure for closing a hierarchy of fluid moment equations is


to obtain the needed closure moments, which are sometimes called constituitive relations, from integrals of the kinetic distribution function f (5.28)
and (5.31) for q and . The distribution function must be solved from a
kinetic equation that takes account of the evolution of the lower order fluid
moments n(x, t), V(x, t), T (x, t) which are the parameters of the lowest order dynamic equilibrium Maxwellian distribution fM specified in (5.22). The
resultant kinetic equation and procedure for determining the distribution function and closure moments are known as the Chapman-Enskog8 approach. In
this approach, kinetic distortions of the distribution function are driven by the
thermodynamic forces T and V gradients of the parameters of the lowest order Maxwellian distribution, the temperature (for q) and the flow velocity
(for ), see (??) in Appendix A.4. For situations where collisional effects are
dominant (/t i << , << 1), the resultant kinetic equation can
be solved asymptotically via an ordering scheme and the closure moments q,
represent the diffusion of heat and momentum induced by the (microscopic) collisions in the medium. This approach is discussed schematically for a collisional
neutral gas in Section A.4. It has been developed in detail for a collisional,
magnetized plasma by Braginskii9 see Section 12.2. While these derivations
of the needed closure relations are beyond the scope of the present discussion,
we will use their results. In the following paragraphs we discuss the physical
processes (phenomenologies) responsible for the generic scaling forms of their
results.
In a Coulomb-collision-dominated plasma the heat flux q induced by a temperature gradient T will be determined by the microscopic (hence the superscript m on ) random walk collisional diffusion process (see Section A.5):
q m T = nT,

m
(x)2

,
n
2t

Fourier heat flux,

(5.48)

in which x is the random spatial step taken by particles in a time t. For


Coulomb collisional processes in an unmagnetized plasma, x (collision
length) and t 1/ (collision time); hence, the scaling of the heat diffusivity is
2 = vT2 / T 5/2 /n. [The factor of 2 in the diffusion coefficient is usually
omitted in these scaling relations because the correct numerical coefficients
(headache factors) must be obtained from a kinetic theory.] In a magnetized
plasma this collisional process still happens freely along a magnetic field (as
long as # << 1), but perpendicular to the magnetic field the gyromotion
limits the perpendicular step size x to the gyroradius (. Thus, in a collisional,
8 Chapman

and Cowling, The Mathematical Theory of Non-Uniform Gases (1952).


Braginskii, Transport Processes in a Plasma, in Reviews of Plasma Physics, M.A.
Leontovich, Ed. (Consultants Bureau, New York, 1965), Vol. 1, p. 205.
9 S.I.

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

20

magnetized plasma we have


# T, # 2 ,
q# = n# b

q = n T,

parallel heat conduction,

(2 , perpendicular heat conduction.

(5.49)

and = b
# = b(
b)

B/B.
with b
Here, as usual, # b
The ratio of the perpendicular to parallel heat diffusion is
/# ((/)2 (/c )2 << 1,

(5.50)

which is by definition very small for a magnetized plasma see (??). Thus, in
a magnetized plasma collisional heat diffusion is much smaller across magnetic
field lines than along them, for both electrons and ions. This is of course the
basis of magnetic confinement of plasmas.
Next we compare the relative heat diffusivities of electrons and ions. From
formulas developed in Chapters 2 and 4 we find that for electrons and ions
with approximately the same temperatures, the electron collision frequency is
> 43 >> 1], the collision lengths are comparable
higher [e /i (mi /me )1/2
> 43 >> 1].
(e i ), and the ion gyroradii are larger [(i /(e (mi /me )1/2
Hence, for comparable electron and ion temperatures we have
#e

#i

mi
me

)1/2

>
43 >> 1,

me
mi

)1/2

< 1
43 << 1.

(5.51)

Thus, along magnetic field lines collisions cause electrons to diffuse their heat
much faster than ions but perpendicular to field lines ion heat diffusion is the
dominant process.
Similarly, the viscous stress tensor caused by the random walk collisional
diffusion process in an unmagnetized plasma in the presence of the gradient in
the species flow velocity V is (see Section 12.2)
2m W,

m
(x)2

,
nm
2t

viscous stress tensor.

(5.52)

Here, W is the symmeterized form of the gradient of the species flow velocity:
W

3 1
12
V + (V)T I( V),
2
3

rate of strain tensor,

(5.53)

in which the superscript T indicates the transpose. Like for the heat flux, the
momentum diffusivity coefficient for an unmagnetized, collisional plasma scales
as m /nm 2 . Similarly for a magnetized plasma we have
m
2
# = 2m
# W# , # /nm ,

m
2
= 2m
W , /nm ( .

(5.54)

b
W b)
b
and W are tensor quantiSince the thermodynamic drives W# b(
tites, they are quite complicated, particularly in inhomogeneous magnetic fields
see Section 12.2. Like for heat diffusion, collisional diffusion of momentum
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

21

along magnetic field lines is much faster than across them. Because of the mass
factor in the viscosity coefficient m , for comparable electron and ion temperatures the ion viscosity effects are dominant both parallel and perpendicular to
B:
(
)1/2
(
)3/2
m
m
me
1
me
#e
5
e
<
<

43 << 1, m m
1.3 10 <<<<< 1.
m
mi
i
i
#i
(5.55)
Now that the scalings of the closure moments have been indicated, we
can use (5.45) to estimate the rate at which entropy increases in a collisional
magnetized plasma. The contribution to the entropy production rate ds/dt
from the divergence of the heat flux can be estimated by ( q)/nT
(# 2# T + 2 T )/T (/T )(2 2# + (2 2 )T . Similarly, the estimated rate
2
of entropy increase from the viscous heating is ( : V)/nT (m
# |# V| +
2
2
2
m
| V| )/nT (|# V/vT | + |( V/vT | ). Finally, the rate of entropy
increase due to collisional energy exchange can be esimated from Qi /ni Ti
e (me /mi ) and Qe /ne Te e (me /mi ) + J 2 / e [me /mi + (V#e V#i )2 /vT2 e ].
For many plasmas the gyroradius ( is much smaller than the perpendicular scale
lengths for the temperature and flow gradients; hence, the terms proportional
to the gyroradius are usually negligible compared to the remaining terms. This
is particularly true for electrons since the electron gyroradius is so much smaller
than the ion gyroradius. We will see in the next section that in the small gyroradius approximation the flows are usually small compared to their respective
thermal speeds; hence the flow terms are usually negligibly small except perhaps
for the ion ones. Thus, the rates of electron and ion entropy production for a
collision-dominated magnetized plasma are indicated schematically by
4 2 2
1
1
(
) 5
e # Te 1 e # Ve 12 me
V#e V#i 2
dse
1
1
+ e max
, 1
,
,
<< e , (5.56)
dt
Te
vT e 1 mi
vT e
4 2 2
1
1 1
1 (
) 5
i # Ti 1 i # Vi 12 1 (i Vi 12 me 1/2
dsi
1, 1
1,
+ i max
, 11
<< i . (5.57)
dt
Ti
vT i 1 1 vT i 1
mi
As shown by the final inequalities, these contributions to entropy production are
all small in the small gyroradius and collision-dominated limits in which they
are derived. Hence, the maximum entropy production rates for electrons and
ions are bounded by their respective Coulomb collision frequencies. For more
collisionless situations or plasmas, the condition # << 1 is usually the first
condition to be violated; then, the collisionless plasma behavior along magnetic field lines must be treated kinetically and new closure relations derived.
Even with kinetically-derived closure relations, apparently the entropy production rates for fluidlike electrons and ion species are still approximately bounded
by their respective electron and ion collision frequencies e and i . However,
in truly kinetic situations with important fine-scale features in velocity space
(localized to v /v << 1), the entropy production rate can be much faster
(ds/dt eff /2 ), at least transiently.
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

22

When there is no significant entropy production on the time scale of interest


(e.g., for waves with radian frequency >> ds/dt), entropy is a constant of the
fluid motion. Then, we obtain the adiabatic (in the thermodynamic sense)
equation of state (relation of pressure p and hence temperature T to density n)
for the species:
1
d
p
ds

ln
+0
dt
1 dt n
Here, we have defined

p n,

T n1,

isentropic equation of state.

= (N + 2)/N,

(5.58)
(5.59)

in which N is the number of degrees of freedom (dimensionality of the system).


We have been treating the fully three dimensional case for which N = 3, = 5/3
and 1 = 2/3 see (5.44). Corresponding entropy functionals and equations
of state for one- and two-dimenional systems are explored in Problems 5.11 and
5.12. Other equations of state used in plasma physics are
p n,
p + 0,

V = 0,

T = constant,
T + 0,
n = constant,

isothermal equation of state ( = 1), (5.60)


cold species equation of state,
(5.61)
incompressible species flow ( ). (5.62)

The last equation of state requires some explanation. Setting ds/dt in (5.58) to
zero and using the density equation (5.37), we find
1 dn
1 dp
=
= V
p dt
n dt

V =

1 1 dp
.
p dt

(5.63)

From the last form we see that for the flow will be incompressible
( V = 0), independent of the pressure evolution in the species. Then, the
density equation becomes dn/dt = n/t + V n = n( V) = 0. Hence,
the density is constant in time on the moving fluid element (Lagrangian picture)
for an incompressible flow; however, the density does change in time in an
Eulerian picture due to the advection (via the V n term) of the fluid into
spatial regions with different densities. Since the pressure (or temperature) is
not determined by the incompressible flow equation of state, it still needs to be
solved for separately in this model.
When one of the regular equations of state [(5.58), (5.60),or (5.61)] is used,
it provides a closure relation relating the pressure p or temperature T to the
density n; hence, it replaces the energy or entropy equation for the species.
When the incompresssible flow equation of state (5.62) is used, it just acts as a
constraint condition on the flow; for this case a relevant energy or entropy equation must still be solved to obtain the evolution of the pressure p or temperature
T of the species in terms of its density n and other variables.

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

5.5

23

Two-Fluid Plasma Description

The density, momentum (mom.) and energy or equation of state equations


derived in the preceding section for a given plasma species can be specialized
to a two-fluid set of equations for the electron (qe = e) and ion (qi = Zi e)
species of charged particles in a plasma:
Electron Fluid Moment Equations (de /dt /t + Ve ):
d e ne
dt
d e Ve
mom.: me ne
dt
3 de Te
ne
energy:
2
dt
or eq. of state: Te
density:

= ne ( Ve )

ne
+ ne Ve = 0, (5.64)
t

= ne e [E + Ve B] pe e + Re ,

(5.65)

= ne Te ( Ve ) qe e : Ve + Qe ,

(5.66)

n1
e .

(5.67)

Ion Fluid Moment Equations (di /dt /t + Vi ):


d i ni
dt
di Vi
mom., mi ni
dt
3 di Ti
ni
energy,
2
dt
or eq. of state: Ti

density,

= ni ( Vi )

ni
+ ni Vi = 0, (5.68)
t

= ni Zi e [E + Vi B] pi i + Ri ,

(5.69)

= ni Ti ( Vi ) qi i : Vi + Qi ,

(5.70)

n1
.
i

(5.71)

The physics content of the two-fluid moment equations is briefly as follows.


The first forms of (5.64) and (5.68) show that in the (Lagrangian) frame of the
moving fluid element the electron and ion densities increase or decrease according to whether their respective flows are compressing ( V < 0) or expanding
( V > 0). The second forms of the density equations can also be written as
n/t|x = V n n V using the vector identity (??); thus, at a given
(Eulerian) point in the fluid, in addition to the effect of the compression or
expansion of the flows, the density advection10 by the flow velocity V increases
the local density if the flow into the local region is from a higher density region
(V n > 0). Density increases by advection and compression are illustrated
in Fig. 5.2. In the force balance (momentum) equations (5.65) and (5.69) the
inertial forces on the electron and ion fluid elements (on the left) are balanced
by the sum of the forces on the fluid element (on the right) Lorentz force
density (nq[E + VB]), that due to the expansive isotropic pressure (p)
and anisotropic stress in the fluid ( ), and finally the frictional force density due to Coulomb collisional relaxation of flow relative to the other species
(R). Finally, (5.66) and (5.70) show that temperatures of electrons and ions
increase due to compressional work ( V < 0) by their respective flows, the
net (divergence of the) heat flux into the local fluid element ( q), viscous
10 See

footnote at bottom of page 15.

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

24

Figure 5.2: The species density n can increase due to: a) advection of a fluid
element by flow velocity V from a higher to a locally lower density region, or
b) compression by the flow velocity V.
dissipation ( : V) and collisional heating (Q) from the other species. Alternatively, when appropriate, the electron or ion temperature can be obtained
from an equation of state: isentropic ( = 5/3), isothermal ( = 1) or cold
species (T + 0).
As written, the two-fluid moment description of a plasma is exact. However,
the equations are incomplete until we specify the collisional moments R and Q,
and the closure moments q and . Neglecting the usually small temperature
gradient effects, the collisional moments are, from Section 2.3:
Electrons:
Ions:

Re + me ne e (Ve Vi ) = ne eJ/, Qe + J 2 / Qi , (5.72)


me
Ri = Re , Qi = 3
e ne (Te Ti ).
(5.73)
mi

For an unmagnetized plasma, the electrical conductivity is the Spitzer electrical conductivity Sp defined in (??) and (??). In a magnetized plasma the
electrical conductivity is different along and perpendicular to the magnetic field.
The general frictional force R and Qe for a magnetized plasma is written as
)
(
J#2
J#
J
J2
R = nq
+
, Qe =
+ Qi , magnetized plasma, (5.74)
#

= (B J/B 2 )B,
in which nq is ne e (electrons) or ni Zi e = ne e (ions), J# J# b
= b(
bJ),

J J J# b
# Sp and 0 . Here, 0 is the reference
electrical conductivity which is defined in (??): 0 ne e2 /me e = 1/, where
is the plasma resistivity.
The closure moments q and are calculated from moments of the distribution function as indicated in (5.28) and (5.31). The distribution function f must
be determined from an appropriate kinetic theory. The closure moments can be
calculated rigorously for only a few special types of plasmas, such as for plasmas where Coulomb collision effects dominate (/t i << , << 1 in
general together with << c , ( << 1 for magnetized plasmas) see Section 12.2. Then, they represent the diffusive transport processes induced by the
(microscopic) Coulomb collision processes in a plasma. For such a plasma the
parametric dependences of the closure moments q, on the collision frequency
and length , and gyroradius ( are indicated in (5.48)(5.55) above for both
unmagnetized and magnetized plasmas.
We will now illustrate some of the wide range of phenomena that are included
in the two-fluid model by using these equations to derive various fundamental
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

25

Figure 5.3: Density distributions of electrons and ions in adiabatic response to


a potential (x).
plasma responses to perturbations. The procedure we will use is to identify the
relevant equation for the desired response, discuss the approximations used to
simplify it and then finally use the reduced form to obtain the desired response.
Since most of these phenomena can occur for either species of charged particles
in a plasma, the species subscript is omitted in most of this discussion.
We begin by considering unmagnetized (B = 0) plasmas. First, consider
the Boltzmann relation adiabatic response (??) to an electrostatic perturbation, which was used in deriving Debye shielding in Section 1.1. It can
be obtained from the momentum equation (5.40), (5.65) or (5.69). Physically an adiabatic description is valid when the thermal motion (pressure in
the two-fluid model) is rapid compared to temporal evolution and dissipative
processes , << vT /x kvT in the language of Section 1.6. Dividing the momentum equation by mnvT and assuming for scaling purposes that
|V| vT , d/dt i, q T, || 1/x k, its various terms are found
to scale as (inertia), kvT (E = electrostatic field force), kvT (pressure
force), (k)2 (stress force), and (frictional force). Thus, for , << kvT (adiabatic regime) and k << 1 (collisional species), the lowest order momentum
equation is obtained by neglecting the inertial force (mn dV/dt) and dissipative
forces due to viscous stress ( ) and collisional friction (R):
0 = nq p.

(5.75)

If we assume an isothermal species [ = 1 in (5.60), (5.67) or (5.71)], the temperature is constant and hence p = T n. Then, we can write the adiabatic force
density balance equation in the form [(q/T ) + ln n] = 0, which in complete
and perturbed form yields
n(x) = n0 eq(x)/T0 ,

q
= ,
n0
T0

isothermal adiabatic response,

(5.76)

This is the usual Boltzmann relation: (??), (??) or (??). As indicated in Fig. 5.3,
in an adiabatic response a potential (x) causes the electron (qe = e < 0)
density to peak where the potential is highest and the ion (qi = Zi e > 0)
density to be at its minimum there. Thus, for an adiabatic response a potential
hill confines electrons but repels ions, whereas a potential valley confines ions
but repels electrons. The adiabatic response for a general isentopic equation
of state [(5.58), (5.67) or (5.71)] is somehat different, although the perturbed
response is the same as (5.76) with the temperature changed to T0 see
Problem 5.13. In addition, the density equation [(5.37), (5.64) or (5.68)] shows
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

26

that perturbed flows are nearly incompressible ( V + 0) in the (adiabatic)


limit of slow changes.
Next, we consider the inertial response, which in the two-fluid context is
usually called the fluid response. It is obtained from a combination of the density and momentum equations. Physically, an inertial response obtains for fast
(short time scale) processes ( >> vT /x kvT ) for which the response to
forces is limited by the inertial force nm dV/dt. Using the same ordering of
the contributions to the momentum equation as in the preceding paragraph,
> kvT >> , the lowest order perturbed (linearized) mobut now assuming

mentum equation becomes mn0 V/t


= n0 q
p. For a plasma species
with a spatially homogeneous density (i.e., n0 = 0), the perturbed density
Thus, in the
equation [(5.37), (5.64) or (5.68)] becomes n
/t = n0 V.
dissipationless, inertial (fluid) limit the density and momentum equations for a
homogeneous plasma species become
n

= n0 V,
t

mn0

V
= n0 q
p.
t

(5.77)

These equations can be combined into a single density response equation by


taking the partial time derivative of the density equation and substituting in
the perturbed momentum equation to yield

n0 q 2 1 2

V
2n
= n0
=
+ p,
2
t
t
m
m

inertial (fluid) response.

(5.78)

The potential fluctuation term represents the inertial polarization charge density
= #0 2 E.
The
derived earlier in (??): 2 pol /t2 = (n0 q 2 /m) E
p
second term on the right of (5.77) represents the modification of this polarization
response due to the thermal motion (pressure) of the species see Problem 5.15.
Alternatively, if we neglect the polarization response, and use a general equation
/t2 (p0 /n0 )2 n
=
of state [(5.58), (5.67) or (5.71)], then (5.77) becomes 2 n
0 which represents a sound wave with a sound wave speed cS (p0 /n0 )1/2
see (??), (??). Note that in the inertial (fluid) limit the perturbed density
0= 0).
response is due to the compressibility of the perturbed flow ( V
We next consider plasma transport processes in a collision-dominated limit.
Specifically, we consider the electron momentum equation (5.65) in a limit where
the electric field force is balanced by the frictional force (R) and the pressure
force:
0 = ne eE pe me ne e (Ve Vi ) = ne eE pe + ne eJ/.

(5.79)

Here, we have neglected the inertia and viscous stress in the collisional limit by
assuming d/dt i << e and e << 1. In a cold electron limit (Te 0)
the last form of this equation becomes
J = E,

Ohms law.

(5.80)

Neglecting the ion flow Vi and using an isothermal equation of state [ = 1 in


(5.67)], we can obtain the electron particle flux (units of #/m2 s) from the first
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

27

Figure 5.4: Unmagnetized plasma particle flux components due to electron diffusion (De ) and mobility (for E = ).
form in (5.79):
Te
e
, M
,
e
me e
me e
electron diffusion, mobility particle fluxes.

e ne Ve = De ne + M
e ne E,

De

(5.81)

The first term represents the particle flux due to the density gradient which is
in the form of a Ficks law (??) with a diffusion coefficient De Te /me e =
vT2 e /2e = e 2e /2. The contribution to the particle flux induced by the electric
field is known as the mobility flux (superscript M ). The directions of these diffusive and mobility particle flux components for an equilibrium (e + 0) electron
species are shown in Fig. 5.4. Note that the electron collision length e = vT e /e
must be small compared to the gradient scale length (i.e., |e ln ne | << 1) for
this collisional plasma analysis to be valid. In general, the ratio of the diffusion coefficient to the mobility coefficient is known as the Einstein relation:
D/M = T /q De /M
e = Te /e. The Einstein relation is valid for many
types of collisional random walk processes besides Coulomb collisions.
Finally, we consider the transport properties embodied in the energy equation for an unmagnetized plasma. Neglecting flows and temperature equilibration between species, the energy equation [(5.43), (5.66) or (5.70)] becomes
3 T
n
= q,
2 t

T
2
= 2 T,
t
3

temperature diffusion.

(5.82)

Here, in the second form we have used the general Fourier heat flux closure relation (5.48) and for simplicity assumed that the species density and diffusivity
are constant in space (n = 0, = 0). In a single dimension this equation
becomes a one-dimensional diffusion equation (??) for the temperature T with
diffusion coefficient D = 2/3 2 . Diffusion equations relax gradients in the
species parameter operated on by the diffusion equation here the temperature gradient for which LT is the temperature gradient scale length defined by
1/LT (1/T )|dT /dx|. From (??) or (5.82) in the form T / T /L2T we infer
that the transport time scale on which a temperature gradient in a collisional
plasma ( << LT ) will be relaxed is (LT /)2 / >> 1/.
As we have seen, the two-fluid equations can be used to describe responses
in both the adiabatic ( << kvT ) and inertial ( >> kvT ) limits. In between,
where kvT , neither of these limits apply and in general we must use a kinetic
equation to describe the responses. Also, we have illustrated the responses for
a collisional species. When Coulomb collision lengths become of order or longer
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

28

Figure 5.5: Flow components in a magnetized plasma.


> 1), the heat flux and viscous stress can no
than the gradient scale lengths (
longer be neglected. However, simultaneously the conditions for the derivation
> 1 we usually need to
of these closure relations break down. Thus, for
use a kinetic equation or theory at least to derive new forms for the closure
relations.

5.6

Two-Fluid Magnetized-Plasma Properties*

We next explore the natural responses of a magnetized plasma using the twofluid model. Because the magnetic field causes much different particle motions
along and across it, the responses parallel and perpendicular to magnetic field
lines are different and must be examined separately. The equation for the evo V is obtained by taking the dot product
lution of the parallel flow V# b
B/B and using
of the momentum equation [(5.40), (5.65) or (5.69)] with b

dV/dt = dV# /dt V db/dt:


b
mn

dV#
R# + mnV db .
= nqE# # p b
dt
dt

(5.83)

Here, the parallel (1) subscript indicates the component parallel to the magnetic
E, # p b
p, R# b
R = nqJ# /# . The responses
field: i.e., E# b
along the magnetic field are mostly just one-dimensional (parallel direction)
forms of the responses we derived for unmagnetized plasmas. However, many
plasmas of practical interest are relatively collisionless along magnetic field
> 1); for them appropriate parallel stress tensor and heat flux
lines (# ln B
closure relations must be derived and taken into account, or else a kinetic description needs to be used for the parallel responses. [See the discussion in the
paragraphs after (??) and (??) in Section 6.1 for an example: the effects of
neoclassical closures for axisymmetric toroidal magnetic systems.]
When the magnetic field is included in the momentum equation [(5.40),
(5.65) or (5.69)], the nqVB term it adds scales (by dividing by mnvT ) to be
of order c ; hence, it is the largest term in the equation for a magnetized plasma
in which c >> , , kvT . Thus, like for the determination of the perpendicular
guiding center drifts in Section 4.4*, the perpendicular flow responses are obtained by taking the cross product of the momentum equation [(5.40), (5.65) or
(5.69)] with the magnetic field B. Adding the resultant perpendicular flows to
the parallel flow, the total flow can be written (see Fig. 5.5)
V = V# + # V + #2 V ,
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

with

(5.84)

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID


(B V)B ,
V # = V# b
B2
EB Bp
V = VE + V
+
,
B2
nqB 2
Bmn dV/dt RB B
+
+
.
V = Vp + V + V
nqB 2
nqB 2
nqB 2

29
(5.85)
(5.86)
(5.87)

Here, the # indicates the ordering of the various flow components in terms of the
small gyroradius expansion parameter # ( (, )/c << 1 see (??)
and (??). As indicated, the cross (subscript ) flow is first order in the small
gyroradius expansion, while the perpendicular (subscript ) flow is second
order compared to the thermal speed vT of the species. For example,
V
Bp
T /m p
( ln p # << 1.
=

vT
nqB 2 vT
(qB/m)vT p

(5.88)

For the scaling of the other contributions to V and V , see Problems 5.19 and
5.20.
The first order flow V VE + V is composed of EB and diamagnetic
flows. The very important EB flow is the result of all the particles in a given
species drifting with the same EB drift velocity (??):
VE

EB
B2

E=

B
1 d

ey ,
+
B2
B0 dx

EB flow velocity. (5.89)

Here and below, the approximate equality indicates evaluation in the sheared
ez and for which plasma parameters (and
slab model of Section 3.1 with B + B0
the potential ) only vary in the x direction. The diamagnetic flow V is
(
)
Bp
T
T (eV)
1 dp
+

ey =

ey ,
V
nqB 2
qB0 p dx
(q/e)B0 Lp
diamagnetic flow velocity,
(5.90)
in which
Lp p/(dp/dx),

pressure-gradient scale length,

(5.91)

which is typically approximately equal to the plasma radius in a cylindrical


model. (The definition of the pressure gradient scale length has a minus sign
in it because the plasma pressure usually decreases with radius or x for a confined plasma.) The last form in (5.90) gives a formula for numerical evaluation
(in SI units, except for T in eV). The V flow is called the diamagnetic flow
because the current density nqV it produces causes a magnetic field that reduces the magnetic field strength in proportion to the species pressure p [see
Problem 5.??], which is a diamagnetic effect. Note that the diamagnetic flows
of electrons and ions are comparable in magnitude and in opposite directions.
Hoewever, the electrical current densities they produce are in the same direction. These diamagnetic currents in the cross (
ey in slab model) direction cause
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

30

charge buildups and polarization of the plasma, which are very important in
inhomogeneous magnetized plasmas.
Of particular importance is the electron diamagnetic flow obtained from
(5.90) with q = e:
(
)
Bpe
Te
Te (eV)
1 dpe
+

ey =

ey ,
Ve =
ne eB 2
eB0 pe dx
B0 Lpe
electron diamagnetic flow velocity.
(5.92)
This is a fundamental flow in a plasma; flows in a plasma are usually quoted
relative to its direction.
The EB and diamagnetic flows are called cross flows because they flow
in a direction given by the cross product of the magnetic field and the radial
gradients of plasma quantities. Thus, they flow in what tends to be the ignorable
coordinate direction the
ey direction in the sheared slab model, the azimuthal
direction in a cylindrical model, or perpendicular to B but within magnetic
flux surfaces in mirror and toroidal magnetic field systems. Since they have no
component in the direction of the electric field and pressure gradient forces (i.e.,
VE E = 0 and V p = 0), they do no work and hence produce no increase
in internal energy of the plasma [i.e., no contributions to (5.41) or (5.42)].
The presence of the EB and diamagnetic flows in a plasma introduces two
important natural frequencies for waves in an inhomogeneneous plasma:
( )
ky T d q
ky d
+
,
EB frequency,
(5.93)
E k VE +
B0 dx
qB0 dx T
(
)
ky T 1 dp
vT
= ky (
, diamagnetic frequency. (5.94)
k V +
qB0 p dx
2Lp
The last approximate form of E is for T = constant.0In the last form of
we have used the definitions of the thermal speed vT 2T /m and gyroradius
( vT /c (??). The electron diamagnetic frequency is often written as
(
)
ky Te 1 dpe
cS
ky Te (eV)
= ky (S
=
,
e +
eB0 pe dx
Lpe
B0 Lpe
electron diamagnetic frequency.
(5.95)
0
in which cS Te /mi is the ion acoustic speed (??) and (S cS /ci .
The significance of the EB frequency is that it is the Doppler shift frequency for waves propagating in the cross direction in a plasma. The significance
of the electron diamagnetic frequency is that it is the natural frequency for an
important class of waves in inhomogeneous plasmas called drift waves (see Section 7.6). Both electron and ion diamagnetic frequency drift waves can become
unstable for a wide variety of plasma conditions (see Section 23.3). Because
drift wave instabilities tend to be ubiquitous in inhomogeneneous plasmas, they
are often called universal instabilities. The presence of the ky ( factor in the
diamagnetic frequencies highlights the significance for drift waves of finite gyroradius effects, mostly due to the ions see (??)(??). The maximum frequency
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

31

Figure 5.6: The diamagnetic flow velocity V can be interpreted physically as


due to either: a) a net
ey flow due to the inhomogeneous distribution of guiding
centers because p = p(x), or b) the combination of the particle guiding center
drifts and magnetization current due to the magnetic moments of the entire
species.
of drift waves is usually limited by finiteion gyroradius effects. For example,
for electron drift waves max{} + vT i /(4 Lpi ) for Te = Ti (see Section 8.6).
Figure 5.6 illustrates two different physical interpretations of the diamagnetic flow. In the fluid, gyromotion picture shown in Fig. 5.6a, because the
density of guiding centers decreases as the radial variable x increases, in a full
distribution of ions executing their gyromotion orbits, more ions are moving
downward (
ey direction) than upward at any given x; hence, dp/dx < 0 in a
ez induces an ion diamagnetic flow in the
ey
magnetized plasma with B + B0
direction see Problem 5.23. In the particle picture shown in Fig. 5.6b, the
flow is produced by a combination of the particle drifts in the inhomogeneneous
magnetic field and the magnetization current due to the magnetic moments of
the charged particles gyrating in the magnetic field, both integrated over the
entire distribution of particles in the species. The electrical current induced by
from (??) and (??),
the guiding center drift velocity dxg /dt = vD vD +vD# b
integrated over an isotropic Maxwellian distribution function fM of particles is
"
EB Bp ( ln B + )
p
(5.96)
bb).
nq
vD q d3v vD f = nq
+
+ b(
B2
B2
B
Here,
used (??)
integrals:
# in evaluating the two types of velocity-space
#
# 3 we have
2
/2)fM = d3v (m/2)(vx2 + vy2 )fM = nT = p and d3v mv#2 fM =
d v (mv
nT = p. The (macroscopic) magnetization due to an entire species of particles
with magnetic moments defined in (??) is given by
"
"
2
mv
M = p b.

bf
(5.97)
M = d3v fM = d3v
2B
B
The electrical current caused by such a magnetization is
JM = M =

Bp Bp ( ln B + )
p

b b).
+
b(
2
2
B
B
B

(5.98)

= b(
b

+ b,

Here, we have used the vector identity (??) and b


b)
into its parallel and perpendicular (to
which can be proved by splitting b
B) components using (??)(??). Comparing these various current components,
we find
vD + M.
(5.99)
nq(VE + V ) = nq
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

32

Thus, for a single species of charged particles in a magnetized plasma, the (fluid,
gyromotion picture) cross () current induced by the sum of the EB and diamagnetic (V ) flows is equal to the (drift picture) sum of the currents induced
by the guiding center particle drifts and the magnetization induced by the magnetic moments of all the particles in the species. Note that no single particle
has a drift velocity that corresponds in any direct way to the diamagnetic flow
velocity V .11 Rather, the diamagnetic flow velocity is a macroscopic flow of an
entire species of particles that is a consequence of the (radially) inhomogeneous
distribution of charged particles in a magnetized plasma. Finally, note that the
net flow of current of a species in or out of an infinitesimal volume does not
vD since M = 0.
involve the magnetization: nq(VE + V ) = nq
Thus, the net flow of (divergence of the) currrent can be calculated from either
the fluid or particle picture, whichever is more convenient.
Next, we discuss the components of the second order perpendicular flow
velocity V Vp + V + V defined in (5.87). The polarization flow Vp
represents the effect of the polarization drifts (??) of an entire species of particles
and to lowest order in # is given by:
(
)
Bmn dV /dt
1 1 d T 1 dp
+

ex ,
+
Vp =
nqB 2
c t B0 dx
q p dx
polarization flow velocity.
(5.100)
Simlarly, we use the first order perpendicular flow V in evaluating the
frictional-force-induced flow V due to the perpendicular component of the frictional force R defined in (5.74):
V

RB
BJ
B[ne e(Ve Vi )]
ne eB(Ve Vi )
=
+
=
2
2
2
nqB
B
B
B 2
(
)
2
(pe + pi )
e (e (pe + pi )
Te + Ti 1 dne
2
=
=
+ e (e

ex ,
0 B 2
2
ne Te
2Ti
ne dx
classical transport flow velocity. (5.101)
=

Here, for simplicity in the evaluation for the sheared slab model form we have
assumed that the electron and ion temperatures are uniform in space and only
the density varies spatially (in the x direction in the sheared slab model). This
flow velocity is in the form of a Ficks diffusion law (??) particle flux

nV = D ne , classical particle flux,


(5.102)
(
)
(
)
Te + Ti
ne Zi
ln
e (2e
m2 /s,(5.103)
+ 5.6 1022 2
2Te
17
B [Te (eV )]1/2

This is called classical transport because its random walk diffusion process
results from and scales with the (electron) gyroradius: x (e . The scaling
11 Many plasma physics books and articles call V the diamagnetic drift velocity. This

nomenclature is very unfortunate since no particles drift with this velocity. Throughout
this book we will call V the diamagnetic flow velocity to avoid confusion about its origin.

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

33

of the particle diffusion coefficient D with collision frequency and gyroradius


is the same as that for the perpendicular electron heat diffusion coefficient e
see (5.49). The particle flux in (5.103) leads to a particle density equation
of the form ne /t = ne Ve = D 2 ne and hence to perpendicular (to
B) diffusion of particles see (5.82), Fig. 5.4 and (??). It is important to
note that the particle flux (and consequent transport) is the same for either
species of particles (electrons or ions). Therefore, it induces no net charge
flow perpendicular to magnetic field lines; hence, it is often said that classical
transport is intrinsically ambipolar electrons and ions diffuse together and
induce no polarization or charge buildup perpendicular to B.
The final perpendicular flow component is:
V

B
,
nqB 2

viscous-stress-induced flow velocity.

(5.104)

For a collisional, magnetized species (# << 1, ( << 1), this flow is


smaller than the classical transport flow velocity V . However, in more col> 1 this flow represents neoclassical transport
lisionless plasmas where #
due to the effects of particles drifting radially off magnetic flux surfaces and it
can be larger than classical transport. For example, for an axisymmetric, large
aspect ratio tokamak, collisions of particles on banana dift orbits (see Section
4.8*) induce a radial particle flux similar to (5.103) with Dr e (2e q 2 #3/2
> 1 is the toroidal winding number of the magnetic field lines and
in which q
# = r/R0 << 1 is the inverse aspect ratio see Chapter 16.
All of the components of the perpendicular flow V have components in
the x or radial (across magnetic flux surface) direction. The polarization flow
leads to a radial current in the plasma and hence to radial charge buildup and
polarization. Because it is due to an inertial force, it is reversible. The radial
flows induced by the frictional and viscous stress forces are due to (microscopic)
collisions and hence yield entropy-producing radial transport fluxes that tend
to relax the plasma toward a (homogeneous) thermodynamic equilibrium.
Finally, it is important to note that like the species flow velocity V, the heat
flow q and stress tensor have similarly ordered parallel, cross (diamagnetictype) and perpendicular components:
q = q# + # q + #2 q ,

= # + # + # ,
2

total conductive heat flux,

(5.105)

total stress tensor.

(5.106)

The scalings of the parallel and pependicular fluxes q# , q and # , with


collision frequency and gyroradius are indicated in (5.49) and (5.54). The cross
heat flux is
q q =

5 nT BT

n bT,
2
qB 2

5 T
,
2 qB

diamagnetic heat flux.

(5.107)
Like the diamagnetic flow, this cross heat flux produces no dissipation [see (5.47)]
since q T = 0. Similarly, the cross stress tensor is a diamagnetic-type tensor
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

34

Table 5.1: Phenomena, Models For An Unmagnetized Plasma


Time,
Length Scales

Species,
Plasma Model

plasma oscillations
Debye shielding

1/pe 1011 s
D 105 m

inertial
adiabatic

q + 0, < pe
q + 0, kD < 1

cold plasma waves

/k > vT ,
/k + cS > vT i

two-fluid
(T + 0, = 0)

oscillations,
dielectric const.

hot plasma waves


Landau damping
velocity-space inst.

< p
< Im{}
< p
< Im{}
< p
< Im{}

Vlasov
Vlasov
Vlasov

dielectric const.
wave damping
NL, via collisions

Coulomb collisions
frequency
length

, k 1
1/ 107 s
0.1 m

plasma
kinetic
equation

two-fluid model

plasma transport

(L/)2 /

two-fluid

loss of plasma

Physical Process

Consequences

of the form mnm


bV and produces no dissipation [see (5.47)] since
: V = 0 see Section 12.2. The cross stress tensor is often called
the gyroviscous stress tensor. Since the gyroviscous effects are comparable to
those from V and q , must be retained in the momentum equations when
diamagnetic flow effects are investigated using the two-fluid equations.

5.7

Which Plasma Description To Use When?

In this section we discuss which types of plasma descriptions are used for describing various types of plasma processes. This discussion also serves as an
introduction to most of the subjects that will be covered in the remainder of
the book. The basic logic is that the fastest, finest scale processes require kinetic descriptions, but then over longer time and length scales more fluidlike,
macroscopic models become appropriate. Also, the equilibrium of the faster
time scale processes often provide constraint conditions for the longer time scale,
more macroscopic processes.
We begin by discussing the models used to describe an unmagnetized plasma.
For specific parameters we consider a plasma-processing-type plasma with Te =
3 eV, ne = 1018 m3 and singly-charged ions (Zi = 1). An outline of the
characteristic phenomena, order of magnitude of relevant time and length scales,
and models used to describe unmagnetized plasmas is shown in Table 5.1. As
indicated in the table, the fastest time scale plasma phenomenon is oscillation
at the electron plasma frequency (Section 1.3) which is modeled with an inertial
DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

35

electron response (5.78). The shortest length scale plasma process is Debye
shielding (Section 1.1), which is produced by an adiabatic response (5.76).
Cold plasma waves (electon plasma and ion acoustic waves) are modeled by
the two-fluid equations by neglecting collisional effects and considering thermal
effects to be small and representable by fluid moments. These natural oscillations result from the dielectric medium responses of the plasma see Chapters
1 and 7. The corresponding hot plasma (kinetic) waves and dielectric functions,
which include wave-particle interaction effects, are modeled with the Vlasov
equation (5.21) and discussed in Chapter 8. Consequences of this kinetic model
of an unmagnetized plasma include the phenomena of collisionless Landau
damping (Section 8.2) of waves and velocity-space instabilities (Chapter 19).
The use of the Vlasov equation is justified because the natural growth or damping rates [Im{}] for these phenomena are larger than the effective collision
frequency. However, velocity-space diffusion due to collisions is required for irreversibility of the wave-particle interactions involved in Landau damping (see
Section 10.2) and to produce a steady state saturation or bounded cyclic behavior during the nonlinear (NL) evolution of velocity-space instabilities (see
Sections 10.3, 24.1, 25.1).
< eff ), Coulomb collisions become important and
On longer time scales (
are modeled using the plasma kinetic equation (5.13). Finally, on transport
time scales (L/)2 / (see Section A.5) long compared to the collision time
1/ and length scales L long compared to the collision length = vT /, the
electron and ion species can be described by the two-fluid equations (5.64)
(5.71). Plasma radiation (caused by particle acceleration via Coulomb collisions
or from atomic line radiation see Chapter 14) can also beome relevant on the
plasma transport time scale. Modeling of plasma particle and energy transport
in collisional plasmas is discussed in Section 17.1.
A similar table and discussion of the relevant phenomena and plasma descriptions on various time and length scales for magnetized plasmas is deferred
to Section 6.8 in the following chapter after we have discussed the important
fast time scale physical effects in a MHD description of a plasma, and in particular Alfv`en waves.
REFERENCES AND SUGGESTED READING
Plasma physics books that provide discussions of various plasma descriptions are
Schmidt, Physics of High Temperature Plasmas (1966,1979), Chapts. 3,4 [?]
Krall and Trivelpiece, Principles of Plasma Physics (1973), Chapts. 2.3 [?]
Nicholson, Introduction to Plasma Theory (1983), Chapts. 3-8 [?]
Sturrock, Plasma Physics, An Introduction to the Theory of Astrophysical, Geophysical & Laboratory Plasmas (1994), Chapts. 11,12 [?]
Hazeltine and Waelbroeck, The Framework of Plasma Physics (1998), Chapts.
36 [?]
Plasma books that provide extensive discussions of plasma kinetic theory are
Klimontovich, The Statistical Theory of Non-equilibrium Processes in a Plasma
(1967) [?]

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

36

Montgomery and Tidman, Plasma Kinetic Theory (1964) [?]


Montgomery, Theory of the Unmagnetized Plasma (1971) [?]
A comprehensive development of the fluid moment equations is given in
S.I. Braginskii,Transport Processes in a Plasma, in Reviews of Plasma Physics,
M.A. Leontovich, Ed. (Consultants Bureau, New York, 1965), Vol. 1, p. 205 [?]

PROBLEMS
5.1 In the year 2000, single computer processor units (CPUs) were capable of about
109 floating point operations per second (FLOPs). Assume a particle pushing code needs about 100 FLOPs to advance a single particle a plasma period
(1/pe ) and that the CPU time scales linearly with the number of particles. How
long would a year 2000 CPU have to run to simulate 0.03 m3 of plasma with
a density of ne = 3 1018 m3 for 103 seconds by advancing all the particles
in a plasma? Taking account of Moores (empirical) law which says that CPU
speeds double every 18 months, how long will it be before such a simulation can
be performed in a reasonable time say one day on a single CPU? Do you
expect such plasma simulations to be possible in your lifetime? /
5.2 Consider a continuum (mush) limit of the plasma kinetic equation. In this
limit charged particles in a plasma are split in two and distributed randomly
while keeping the charge density, mass density and species pressure constant.
Then, the particles are split in two again, and the splitting process repeated an
infinite number of times. What are the charge, mass, density and temperature
of particles in one such split generation relative to the previous one? Show that
in this limiting process the plasma frequency and Debye length are unchanged
but that the term on the right of the averaged Klimontovich equation (5.12)
becomes negligibly small compared to the terms on the left. Use these results to
discuss the role of particle discreteness versus continuum effects in the Vlasov
equation and the plasma kinetic equation. //
5.3 Show that for a Lorentz collision model the right side of the averaged Klimontovich equation (5.12) becomes the Lorentz collision operator:
"v v# f

v
2t
v
in which "vv#/t is given by (??). [Hint: First subtract the averaged
Klimontovich equation (5.12) from the full Klimontovich equation (5.8) and show
m
that df m
= (q/m)
average
# /dt
# E# f#/v. Then, for an ensemble
# t defined by
3
"g# = ni d x g = ni v dt b db d g show that (q/m)2 "Em dt$ Em # =
"v v#/2t.] ///
CL (f ) =

5.4 Use the Lorentz collision operator defined in the preceding problem to show that
for a Maxwellian distribution with a small flow (|V|/vT << 1) the Coulomb
collision frictional force density on an electron species in the ion rest frame is
Re = me ne e Ve . //
5.5 Show that the partial time derivative of the Maxwellian distribution (5.22) is
fM
=
t

1 n
1 T
+
n t
T T

mvr2
3

2T
2

m
V
+
vr
T
t

fM .

Also, derive similar expressions for fM fM /x and fM /v. //

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

37

5.6 Write down a one-dimensional Vlasov equation governing the distribution function along a magnetic field line neglecting particle drifts. What are the constants
of the motion for this situation? What is the form of the general solution of this
Vlasov equation? Discuss what dependences of the distribution function on the
constants of the motion are needed to represent electrostatic and magnetic field
confinement of the charged particles in a plasma along B. //
5.7 Show the integration and other steps needed to obtain the energy equation
(5.41). [Hint: For the velocity derivative term derive and use the vector identity
mv 2
mv 2
A(v) =

A(v) mv A(v).]
2 v
v
2
Also, use the origin of the energy flux (5/2)nT V to show that it represents a
combination of the convection of the internal energy and mechanical work done
on or by the species moving with a flow velocity V. //*
5.8 Show the steps in going from the first energy equation (5.41) to the second
(5.42). [Hint: Use vector identities (??) and (??).] //*
5.9 Parallel electron heat conduction often limits the electron temperature that can
be obtained in a collisional magnetized plasma that comes into contact with
the axial end walls. a) Develop a formula for estimating the equilibrium central
electron temperature Te (0) produced by a power source supplying QS watts per
unit volume in a plasma of length 2L that loses energy to the end walls primarily
by parallel electron heat conduction. For simplicity, neglect the variation of the
parallel heat conduction with distance & along a magnetic field line and assume
a sinusoidal electron temperature distribution along a magnetic field line given
by Te (&) = Te (0) cos(&/2L). b) How does Te (0) scale with QS ? c) For a plasma
with singly-charged ions and ne = 1012 cm3 in a chamber with an axial length
of 1 m, what Te (0) can be produced by a power source that supplies 0.1 W/cm3
to the plasma electrons? d) How large would QS need to be achieve a Te (0) of
25 eV? //*
5.10 The irreducible minimum level of perpendicular heat transport is set by classical
plasma transport. Consider an infinitely long cylinder of magnetized plasma.
Estimate the minimum radius of a 50% deuterium, 50% tritium fusion plasma
at Te = Ti = 10 keV, ne = 1020 m3 in a 5 T magnetic field that is required to
obtain a plasma energy confinement time of 1 s. //*
5.11 Write down one- and two-dimensional Maxwellian distribution functions. Use
the entropy definition in (5.44) to obtain entropy functionals for these two distributions. Show that the entropy functions are as indicated in (5.58). //*
5.12 First, show that in N dimensions the energy equation (5.42) can be written, in
the absence of dissipative effects, as
N p
=
2 t

N +2
pV + V p.
2

Then, show that in combination with the density equation (5.37) this equation
can be rearranged to yield the isentropic equation of state in (5.58). //*
5.13 Derive the adiabatic response for an isentropic equation of state. Show that the

perturbed adiabatic response is n


/n0 & q /T
0 in which T0 p0 /n0 . //

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

38

5.14 Use the ion fluid equations (5.68)(5.71) to derive the ion energy conservation
relation (??) that was used in the analysis of a plasma sheath in Section 1.2.
Discuss the various approximations needed to obtain this result. //
5.15 Use the inertial electron fluid response (5.78) with a general isentropic equation
of state to obtain the thermal speed corrections to the electron plasma wave
dielectric )I (??). Set the dielectric function to zero and show that the normal
modes of oscillation satisfy the dispersion relation
2
2 = pe
+ (/2) k2 vT2 e . //

5.16 Use the two-fluid equations (5.64)(5.71) to obtain the ion sound wave equation
(??). Also, use the two-fluid equations and an isothermal equation of state for
the ions to obtain the ion thermal corrections to the ion acoustic wave dispersion
relation (??). //
5.17 Show how to use the electron fluid equations to derive the electromagnetic skin
depth defined in (??). /
5.18 Consider a collisional unmagnetized plasma where the electron density distribution ne (x) is determined by some external means, for example by a combination
of wave heating and ionization of neutrals. Use the equilibrium Ohms law (electron momentum equation) in (5.79) to determine the potential distribution (x)
(for E = ) required to obtain no net current flowing in the plasma. For
simplicity assume isothermal electrons. Then, use this potential to show that
the equilibrium distribution of isothermal ions of charge Zi in this plasma is
ni (x)/ni (0) = [ne (0)/ne (x)]Zi Te /Ti.
What is the role of the potential (x) here? Explain why the ion density is
smallest where the electron density is the largest in this plasma situation. //
5.19 Show that for q T the EB flow is order ) relative to the thermal speed of
the species in the small gyroradius expansion. /
5.20 Show that all the terms in the V defined in (5.87) are of order )2 (or smaller)
relative to the thermal speed of the species in the small gyroradius expansion.
[Hint: Use the first order EB and diamagnetic cross flows to estimate the
various contributions to V .] //
5.21 Suppose a drift-wave has a real frequency of 0.5 i in the EB rest frame and
that ni qi = 2 pi , ky = 0.1 cm1 and d/dx = 100 V/cm with a magnetic field of 2.5 T. What is the frequency (in rad/s and Hz) of the wave in the
laboratory frame? Does the wave propagate in the electron or ion diamagnetic
flow direction in the laboratory frame? /
5.22 Calculate the diamagnetic flow velocity in a uniform magnetic field from a simple kinetic model as follows. First, note that since the relevant constants of the
motion are the guiding center position xg = x + vy /c from (??) and energy
g , an appropriate solution of the Vlasov equation is f = f (xg , g ). Assume
a Maxwellian energy distribution and expand this distribution in a small gyroradius expansion. Show that the flow velocity in this expanded distribution is
the diamagnetic flow velocity (5.90). Discuss how this derivation quantifies the
illustration of the diamagnetic flow in Fig. 5.6a. //

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 5. PLASMA DESCRIPTIONS I: KINETIC, TWO-FLUID

39

5.23 Consider electron and ion pressure profiles peaked about x = 0 in a sheared slab
magnetic field model with no curvature or shear. a) Sketch the directions of
the diamagnetic flows of the electrons and ions. b) Show that the currents they
induce are in the same direction. c) Show that these currents have a diamagnetic
effect on the magnetic field strength. d) Finally show that for each species the
induced diamagnetic change in the magnetic field energy density is proportional
to the pressure of the species. /
5.24 Consider a plasma species with an anistropic Maxwellian-type distribution that
has different temperatures parallel and perpendicular to the magnetic field but
no dependence on the gyrophase angle . a) Show that for this anisotropic
b)
+ p& b
b.
b) Show that for an
distribution the pressure tensor is P = p (I b
anisotropic species the diamagnetic flow velocity is
V

B[p + (p& p )]
B P
=
.
nqB 2
nqB 2

c) Calculate the velocity-space-average drift current nq


vD , magnetization M
and magnetization current JM for an anisotropic species. d) Show that your
results reduce to (5.96)(5.98) for isotropic pressure. e) Finally, show that (5.99)
is also satisfied for a plasma species with an anisotropic pressure. ///
5.25 In the derivation of (5.99) we neglected the guiding center drift due to the

direction of the magnetic field changing in time the b/t


contribution.
D , (5.99) must be modified by adding
Show how, when this drift is included in v
the part of the polarization flow Vp caused by V& to its left side to remain valid.
(Assume for simplicity that the magnetic field is changing in direction slowly

compared to the gyrofrequency [(1/c )| b/t|


<< 1] so the small gyroradius
expansion used to derive the guiding center orbits is valid.) //
5.26 Show that classical diffusion is automatically ambipolar for a plasma with multiple species of ions. [Hint:6Note that because of momentum conservation in
Coulomb collisions Re = i Ri .] //

DRAFT 11:54
January 21, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

Chapter 6

Plasma Descriptions II:


MHD
The preceding chapter discussed the microscopic, kinetic and two-fluid decsriptions of a plasma. But we would actually like a simpler model one that would
include most of the macroscopic properties of a plasma in a one-fluid model.
The simplest such model is magnetohydrodynamics (MHD), which is a combination of a one-fluid (hydrodynamic-type plus Lorentz force effects) model for
the plasma and the Maxwell equations for the electromagnetic fields. The main
equations, properties and applications of the MHD model are developed in this
chapter.
In the first section, we further approximate and combine the two-fluid description in Section 5.5 to obtain a one-fluid magnetohydrodynamics (MHD)
description of a magnetized plasma. Section 6.2 presents the MHD equations in
various forms and discusses their physical content. Subsequent sections discuss
general properies of the MHD model (force-balance) equilibria (Section 6.3),
boundary and shock conditions (Section 6.4), dynamical responses (Section 6.5),
and the Alfv`en waves (Section 6.6) that result from them. Then, Section 6.7 discusses magnetic field diffusion in the presence of a nonvanishing plasma electrical
resistivity. Finally, Section 6.8 discusses the relevant time and length scales on
which the kinetic, two-fluid and MHD models of magnetized plasmas are applicable, and hence usable for describing various magnetized plasma phenomena.
This chapter thus presents the final steps in the procedures and approximations
used to progress from the two-fluid plasma model to a macroscopic description,
and discusses the key properties of the resultant MHD plasma model.

6.1

Magnetohydrodynamics Model*

Magnetohydrodynamics (MHD) is the name given to the nonrelativistic single


fluid model of a magnetized (, i << ci ), small gyroradius (#i << 1)
plasma. The MHD description is derived in this section by adding appropriDRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

ately the two-fluid equations [(??)(??)] to obtain a one-fluid description and


then making suitable approximations. The philosophy of the ideal MHD description is to obtain density, momentum and equation of state equations that
govern the macroscopic behavior of a magnetized plasma on fast time scales
where dissipative processes are negligible and entropy is conserved. Thus, ideal
MHD processes are isentropic. The philosophy of resistive MHD is to extend
the time scale beyond the electron collision time scale ( 1/e ) by adding to
ideal MHD the irreversible, dissipative effects due to the electrical resistivity in
the plasma.
The pedagogical approach we will use is to first define the MHD plasma
variables and next obtain conservation equations for these quantities. Then, we
discuss the approximations used in obtaining the MHD plasma equations, and
finally (in the next Section) we summarize the equations that constitute the
MHD model of a plasma and its electromagnetic fields. We begin by defining
the one-fluid plasma variables of MHD:
!
m
ms ns = me ne + mi ni $ mi ni
(6.1)
mass density (kg/m3 ):
s

mass flow velocity (m/s):

current density (A/m2 ):

"
me ne Ve + mi ni Vi
s ms n s V s
=
$ Vi
V "
m
n
m
s s
s
(6.2)
!
J
ns qs Vs = ne e(Ve Vi )
(6.3)
s

plasma pressure (N/m ):


2

stress tensor (N/m2 ):

n s ms 2 $
|Vs | $ pe + pi
(6.4)
3
s
&
'(
!%
1 2

=
s + ns ms Vs Vs I |Vs |
3
s
P

!#

ps +

$ e + i ,

(6.5)

s Vs V is the species flow velocity relative to the mass flow


in which V
velocity V of the entire plasma. Here, the forms on the right indicate first the
general form as a sum over the species index s, second the electron-ion two-fluid
form, and finally, after an appoximate equality, the usual, approximate forms
< 1/1836 <<< 1, comparable Ve and Vi , and |Vi | << vT i . By
for me /mi
construction, the pressure and stress tensor are defined in the flow velocity rest
frame, which is often called the center-of-mass (really momentum) frame see
Problem 6.1.
A one-fluid mass density (continuity) equation for the plasma is obtained
by multiplying the electron and ion density equations (??) and (??) by their
respective masses to yield m /t + m V = 0. Multiplying the density equations by their respective charges qs and summing over species yields the charge
continuity equation q /t + J = 0. In MHD the plasma is presumed to
be quasineutral because we are interested in plasma behavior on time scales
long compared to the plasma period ( << p ) and length scales long compared to the Debye shielding distance (D /x kD << 1). Mathematically,
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

"
quasineutrality in the plasma means q s ns qs = e(Zi ni ne ) $ 0. Thus,
in the MHD model the charge continuity equation simplifies to J = 0. Note
that this equation is also consistent with the divergence of Amperes law when
the displacement current is neglected see (??). Hence, the charge continuity
equation J = 0 is also consistent with a nonrelativistic MHD description of
particles and waves in a plasma. Since MHD plasmas are quasineutral and have
no net charge density (q = 0), the Gauss law Maxwell equation E = q /(0
cannot be used to determine the electric field in the plasma. Rather, since
a plasma is a highly polarizable medium, in MHD the electric field E is determined self-consistently from Ohms law, Amperes law and the charge continuity
equation ( J = 0).
A one-fluid momentum equation (equation of motion) for a plasma is obtained by simply adding the electron and ion momentum equations (??) and
(??) (see Problem 6.2 for the structure of the inertia term m dV/dt):
m

dV
= q E + JB P ,
dt

(6.6)

in which $ e + i is the total plasma stress tensor in the center-of-mass


frame defined in (6.5). The electric field term is eliminated in MHD by the
assumption of quasineutrality in the plasma: q $ 0. In a collisional plasma
the viscosity effects of the ions are dominant in the stress tensor [see (??)].
The dissipative effects due to ion viscosity become important on time scales
long compared to the relatively slow ion collision time scale [see (??)]. For low
collisionality plasmas in axisymmetric toroidal magnetic systems these parallel
"i ) represent the viscous drag on the parallel
ion viscosity effects (due to b
(poloidal) ion flow carried by untrapped ions due to their collisions with the
stationary trapped ions, and are included in a model called neoclassical MHD;
there they result in damping of the poloidal ion flow at a rate proportional to
the ion collision frequency i and consequently to an increased perpendicular
inertia and dielectric response for t >> 1/i see Chapter 16. In ideal and
resistive MHD it is customary to neglect the viscous stress effects and thus set
= 0 in (6.6). This assumption is usually valid for time scales shorter than
the ion collision time scale: d/dt i >> i .
Since the magnetic field causes the plasma responses to be very different
along and transverse to the magnetic field direction, it is useful to explore the
B/B
responses in different directions separately. Taking the dot product of b
with the plasma momentum equation (6.6) and neglecting q E (quasineutrality
assumption) and the stress tensor , the parallel plasma momentum equation
becomes

dV"
db
= " P m V
.
(6.7)
m
dt
dt
= /). The last term is important only when the magnetic
in which " b
field direction is changing in time or in inhomogeneous plasmas when the flow
velocity V is large. Neglecting this term, (6.7) in combination with the plasma
mass density (continuity) equation leads to compressible flows due to plasma
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

pressure perturbations and hence to sound waves along the magnetic field
see (??)(??) in Section A.6 and (6.89) below.
Taking the cross product of B with the momentum equation and using the
bac cab vector identity (??), again neglecting q E and the stress tensor , we
obtain the two perpendicular components of the current:
J

Jp

BP
,
B2
Bm dV/dt
,
B2

diamagnetic current density,

(6.8)

polarization current density.

(6.9)

The diamagnetic current is the sum of the currents produced by the diamagnetic currents due
" to flows in the various species of charged particles in the
plasma: J = s ns qs Vs . Like the species diamagnetic flows, it is called a
diamagnetic current because it produces a magnetic field that reduces the
magnetic field strength in proportion to the plasma pressure P (see Problem
6.13). The electric field produces no perpendicular current in MHD because
the
of all species are the same; hence, they produce no current:
" EB flows "
n
q
V
=
(
s s s Es
s ns qs )VE = q VE $ 0.
Like for the individual species diamagnetic flows [see (??) and Fig. ??], the
(fluid picture) diamagnetic current is equal to the (particle picture) current due
to the combination of the particle guiding center drifts and the magnetization
produced by the magnetic moments () of all the charged particles gyrating in
the B field:
(6.10)
J = JD + M,
in which the particle drift (D) and the magnetization (M ) currents are
JD
JM

BP ( ln B + ) P

+ b(b b),
B2
B
!)
!
b
P
M
ps = b.
d3v s fM s =
B
B
s
s

Ds =
ns q s v

M,

(6.11)
(6.12)

Note that since the (dimensionless) magnetic susceptibility M is defined by


M = M B/0 [see (??)], in the MHD model of the plasma M = (0 P/B 2 ).
The negative sign of M indicates the diamagnetism effect of the magnetic
moments of the gyrating particles in a magnetized plasma. As an illustration of
the magnitude of this diamagnetism effect, when the plasma pressure P is equal
to the magnetic energy density [see (??)] B 2 /20 , the magnetic field strength
is halved.
The polarization current is the current produced by the sum
" of the currents
due to the polarization flows of the various species: Jp = s ns qs Vp . Since
the ion mass is so much larger than the electron mass, the ion polarization flow
dominates: Jp $ ni Zi eVpi . There is no resistivity-driven current (i.e., no J )
because the classical diffusion induced by the plasma resistivity is ambipolar
[see (??)]. Also, there is no viscosity-induced current (i.e., no J ) in MHD
because the stress tensor effects are neglected, assuming >> i .
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

The total current in MHD is a combination of the parallel current, and the
diamagnetic and polarization perpendicular currents:
J = J" + J + Jp = J"

B BP
Bm dV/dt
+
+
.
B
B2
B2

(6.13)

J =
The parallel component of the current density is defined by J" b
(B J)B. Quasineutrality of the highly polarizable, magnetized plasma is ensured in MHD through
0 = J = (B )(J" /B) + J + Jp ,
MHD charge continuity equation,

(6.14)

which is a very important equation for analyzing MHD equilibria and instabilities. The derivative of the parallel current has been simplified here using the
vector identity (??) and the Maxwell equation B = 0:
J" = (J" /B)B = (B )(J" /B)+(J" /B) B = (B )(J" /B). (6.15)
Taking the divergence of the diamagnetic current equation (6.20), we obtain
(see Problem 6.3)
B( ln B + )
1
b),

P )(b
P + (b
2
B
B
P )(0 J" /B 2 ).
= J ( ln B + ) + (b
(6.16)

J = JD =

Here, we have used vector identities (??) and (??) to evaluate the divergence
b
= 0 J B/B 2 = 0 J" /B see
of J and Amperes law to write b
discussion after (??). Thus, like for the individual species current contributions,
the net (divergence of the) electrical current flow in or out of an infinitesimal
volume can be computed from either the divergence of the diamagnetic current
(fluid picture) or the divergence of the particle drift current (particle picture).
The important effects of the (mostly radial) pressure gradients in the MHD
model of a magnetized plasma are manifested through the diamagnetic current J it induces and, for inhomogeneous magnetic fields, the net charge flows
induced [see (6.16)]. For the MHD charge continuity equation (6.14) to be satisfied, compensating parallel (J" ) or polarization (Jp ) currents must flow in the
plasma. These electrical currents can lead, respectively, to modifications of the
MHD equilibrium (Chapter 20) and pressure-gradient-driven MHD instabilities
(Chapter 21).
Next, we obtain an Ohms law for MHD. A one-fluid generalized Ohms
law is obtained by multiplying the electron and ion momentum equations by
qs /ms and summing them to produce an equation for J/t see Problem 6.4.
However, we proceed more physically and directly from the electron momentum
equation. Using Ve = Vi J/ne e $ V J/ne e and the anisotropic frictional
force R in (??), and dividing the electron momentum equation (??) by ne e,
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD


we find it can be written (to lowest order in me /mi ) as
& '
&
'
J"
J
m e d Je
JB pe e
,
+
= E + VB

e2 dt ne
"

ne e
generalized Ohms law.

(6.17)

Here, we have neglected an ion flow inertia term on the left because it is order
< 1/1836 smaller than the inertial flow contribution coming from the
me /mi
Jp B term evaluated using the polarization current (6.9). While the first and
third terms on the right indicate a simple Ohms law E = J/, there are a
number of additional terms. To understand the role and magnitude of these
other contributions to the generalized Ohms law and obtain an MHD Ohms
law, we need to explore separately their contributions along and perpendicular
to the magnetic field direction.
) of the generalized Ohms law is:
The parallel component (b
de (J/ne )/dt = E" J" /" + (" pe + b
e )/ne e.
(me /e2 ) b

(6.18)

The electron inertia term on the left is small compared to E" for scale lengths
longer than the electromagnetic skin depth (see Section 1.5): |(c/pe )|
kc/pe << 1 see Problem 6.5. Since c/pe is typically a very short distance
(c/pe $ 103 m = 1 mm for ne $ 3 1019 m3 ), this is usually a good approximation in MHD which seeks to provide a plasma description on macroscopic
scale lengths. Also, since 1/" me e /ne e2 , the electron inertia term is of
of order /e compared to the parallel friction force term J" /" . In resistive
MHD it is assumed that << e so the electron inertia can be neglected in the
parallel Ohms law.
The parallel electron pressure gradient term is neglected in MHD because of
a fundamental approximation in MHD that electric field effects are larger than
pressure gradient effects:
|E" | >> |" P |/ne e,

|E | >> | P |/ne e,

MHD approximations. (6.19)

Physically, the MHD model describes situations in which collective electric field
effects are more important than the thermal motion (pressure) effects of both
electrons and ions. Mathematically, this approximation is appropriate (both
along and across magnetic field lines see Problem 6.6) when the EB flow
velocity VE is large compared to the diamagnetic flow velocities Ve , Vi and
hence for , E >> e , i .
Finally, we consider the contribution due to the parallel component of the
viscous stress. While this term is negligible compared to J" /" in a collisional plasma [see (??)], it can be important in more collisionless plasmas where
> 1 in which e = vT e /e is the electron collision length. For low colli e "
sionality plasmas in axisymmetric toroidal magnetic systems these parallel elec "e ) represent the viscous drag on the parallel
tron viscosity effects (from b
electron flow carried by untrapped electrons due to their collisions with the
stationary trapped electrons and ions, and they are included in a model called
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

neoclassical MHD; there they result in order unity modifications of the parallel
Ohms law (see Chapter 16) reductions in the parallel electrical conductivity
and a so-called bootstrap current parallel to B induced by the radial gradient
of the plasma pressure. In ideal and resistive MHD the parallel electron inertia,
pressure gradient and viscosity effects are all neglected and the parallel Ohms
law becomes simply E" = J" /" .
Next, we consider the perpendicular component of the generalized Ohms
b
):
law. It is obtained by operating on (6.17) with b(
0 = E + VB + J / [ JB pe ( e ) ]/ne e

(6.20)

in which the subscript indicates the component perpendicular to B [see (??)].


The perpendicular electron inertia term has been neglected here because it is a
< (1/1836)(/ci ) <<< 1 smaller
factor of at least /ce = (ci /ce )(/ci )
than the E term and hence negligible in MHD see Problem 6.7. The first
two terms on the right give the dominant part of the perpendicular Ohms law
and when set to zero yield a perpendicular plasma flow velocity V = VE =
EB/B 2 . The JB term on the right is known as the Hall term; it indicates
a perpendicular electric field caused by current flowing transverse to a magnetic
field. In MHD the perpendicular current is composed of the diamagnetic and
polarization currents defined in (6.8) and (6.9). The diamagnetic Hall term
component J B = P , and the pe and ( e ) terms are comparable
in magnitude; they are all neglected in MHD because of the perpendicular part
of the MHD approximation (6.19). Finally, the ratio of the polarization current
contribution in the Hall term to the electric field term is |Jp B|/(ne e|E |)
(m /ne e)|dV /dt|/|E | (1/ci )|dE /dt|/|E | /ci , which is small in
the small gyroradius expansion necessary for the validity of MHD. Thus, our
perpendicular Ohms law in MHD becomes simply E + VB = J / .
The perpendicular Ohms law can be combined with the MHD parallel Ohms
law to yield
E + VB = J" /" + J / ,

complete MHD Ohms law.

(6.21)

The parallel electrical conductivity " is at most a factor [see (??)] of 1/e
32/3 $ 3.4 greater than the perpendicular conductivity = 0 . Thus, it
is customary in resistive MHD to not distinguish the electrical conductivity
along and transverse to the magnetic field, but instead to just use an isotropic
electrical resistivity defined by 1/0 = me e /ne e2 . Hence, the MHD Ohms
law is usually written as simply E + VB = J.
In MHD the Ohms law is used to write the electric field in terms of the flow
velocity V and current J. Taking the cross product of the Ohms law with the
magnetic field B, we obtain the perpendicular MHD mass flow velocity V :
V =

EB BJ
+
= VE + V .
B2
B2

(6.22)

Thus, the perpendicular MHD mass flow velocity is the sum of the EB flow
velocity (??) and the (ambipolar) classical transport flow velocity (??), which
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

although small is kept because it is a consequence of including resistivity in


the Ohms law. [The diamagnetic flow velocity V does not appear in the
perpendicular MHD mass flow velocity V because of the MHD approximation
(6.19); the polarization flow Vp and viscosity-driven flow V are not included in
the MHD V because they are higher order in the small gyroradius expansion.]
) component of the MHD Ohms law (??) yields
The parallel (b
E" = J" .

(6.23)

In the ideal MHD limit where 0, this equation requires E" = 0, which
for a general E = A/t is satisfied in equilibrium by the equilibrium
potential being constant along the magnetic field, and in perturbations by the
parallel gradient of the potential being balanced by a parallel inductive (vector
" = " A" /t = 0.
potential) component: E
Finally, we need a one-fluid energy equation or equation of state to close
the hierarchy of MHD equations. In MHD it is customary to use an isentropic
equation of state (d/dt) ln(P/m ) $ 0. Using P = pe +pe , 3/2 = 1/(1) and
working out the time derivative in terms of the time derivatives of the electron
and ion entropies given in (??), (??), (??) and (??), we obtain
&
'
+
P
1
1*
dse
dsi
d
ln
=
pe
+ pi
$
qe Vi : i + J 2 .
dt m
P
dt
dt
P
(6.24)
The last, approximate form indicates the dominant contributions to the overall
plasma entropy production rate. Its last term indicates entropy production by
joule heating; while this rate is usually small [$ e (|J|/ne evT e )2 << e , of order
one over the plasma confinement time], it should be kept in resistive MHD for
consistency with the inclusion of resistivity in the Ohms law. As discussed
after (??), the ion viscous dissipation rate is at most of order the ion collision
frequency i for fluidlike ions; thus, like the ion viscous stress tensor effects
in the plasma momentum equation, it is usually neglected assuming d/dt
i >> i .
Most problematic for an isentropic plasma equation of state is the electron
heat conduction. In a collisional plasma, parallel electron heat conduction leads
to a plasma entropy production rate of order e (e " )2 << e , which is often
smaller than MHD wave frequencies and hence negligible. However, in low
> 1, parallel electron heat conduction can
collisionality plasmas where e "
cause entropy production rates of order e or perhaps larger [see disussion after
(??)], which can be of order MHD wave frequencies. On the other hand, if the
electron fluid responds totally collisionlessly, there is no entropy production from
electron heat conduction (or any other collisionless electron process). In MHD
it is customary to neglect the electron heat conduction contributions to entropy
production on the basis that either: 1) d/dt i >> e ; 2) parallel electron
temperature gradients are quite small because of parallel heat conduction and
thus lead to a negligible entropy production rate [ >> e 2e (2" T )/T ]; or 3) the
relevant electron response is totally collisionless and hence leads to no entropy
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

production. However, there could be circumstances where entropy-producing


parallel electron heat conduction effects are important on MHD wave time scales.

6.2

MHD Equations

The equations used to describe the MHD model of a magnetized plasma and
the associated electric and magnetic fields are thus given by
MHD Plasma Description (Ideal, 0; Resistive, ,= 0):
mass density:
charge continuity:
momentum:
Ohms law:
equation of state:
total time derivative:

m
+ m V = 0,
t
J = 0,
dV
= JB P,
m
dt
E + VB = J,
P
J 2
d
ln = ( 1)
$ 0,
dt m
P

+ V.
dt
t

(6.25)

B
= E,
t
B = 0,
0 J = B.

(6.31)

(6.26)
(6.27)
(6.28)
(6.29)
(6.30)

Maxwell Equations for MHD:


Faradays law:
no magnetic monopoles:
nonrelativistic Amperes law:

(6.32)
(6.33)

Gauss law ( E = q ) does not appear in the list of Maxwell equations because
in the MHD model plasmas are highly polarizable, quasineutral (q $ 0) fluids in
which the electric field is determined self-consistently from Ohms law, Amperes
law and the charge continuity equation J = 0.
The MHD model describes a very wide range of phenomena in small gyroradius, magnetized plasmas macroscopic plasma equilibrium and instabilities,
Alfv`en waves, magnetic field diffusion. It is the fundamental, lowest order model
used in analyzing magnetized plasmas.
The physics content of the MHD plasma description is briefly as follows.
The equation for the mass density (m $ mi ni ) is also called the continuity
equation and can be written in the form m /t = Vm m V. When
written in the latter form, it describes changes in mass density due to advection
(Vm ) and compressibility (V ,= 0) by the mass flow velocity V see
Fig. ??. The charge continuity equation is the quasineutral (q $ 0) form of
the general charge continuity equation q /t + J = 0 that results from
adding equations for the charge densities of the electron and ion species in
the plasma. [While J = 0 also results from taking the divergence of the
nonrelativistic (i.e., without displacement current) Amperes law, it is often
better to think of it as the equation that ensures quasineutrality of the plasma
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

10

in the MHD model as indicated in (6.14).] The momentum equation, which


is also known as the equation of motion, provides the force density balance for
a fluid element (infinitesimal volume of fluid) that is analogous to ma = F for
a particle: the inertial force (m dV/dt) is equal to the magnetic force (JB)
plus the (expansive) pressure gradient force (P , where P = pe + pi is the
total plasma pressure) on a fluid element. The MHD Ohms law, which is a
simplified form of the electron momentum equation, is just the basic laboratory
frame Ohms law E& = J for a fluid moving with plasma mass flow velocity
V: E& = E + VB. The MHD equation of state is an isentropic (adiabatic
in thermodynamics) equation of state except for the small entropy production
rate by joule heating ( J 2 /P 1/E ), which is usually negligibly small but
is retained for consistency with inclusion of resistivity in Ohms law. The total
time derivative in (6.30) indicates that time-differentiated quantities change
both because of local (Eulerian) temporal changes (/t|x ) and because of being
carried along (advected) with the MHD fluid (V) at the velocity V.
After some manipulations, it can be shown (see Problems 6.86.9) that the
MHD equations yield the following conservative forms of total MHD system
mass, momentum and energy relations:
MHD system mass equation:
MHD system momentum equation:
MHD system energy equation:
in which
MHD stress tensor:
MHD energy density:
MHD energy flux:

m
+ m V = 0,
t
(m V)
+ T = 0,
t
w
+ S = 0,
t
&

'
B2
BB
,
T m VV + P +
I
20
0
P
B2
m V 2
+
+
,
w
2
1 20
'
&

EB
m V 2
+
P V+
.
S
2
1
0

(6.34)
(6.35)
(6.36)

(6.37)
(6.38)
(6.39)

Here, the contributions to the MHD system stress tensor are due to the flow
(m VV, Reynolds stress), isotropic pressure (P I) and both isotropic expansion
[(B 2 /20 )I] and tension (BB/0 ) stresses in the magnetic field see (??).
The Reynolds stress is only important in systems with large flow; it is negligible
in MHD systems with strongly subsonic flows (m V 2 /2P V 2 /c2S << 1). The
system energy density is composed of the densities of the kinetic (flow) energy
(m V 2 /2), internal energy (3P/2 for a three-dimensional system with = 5/3)
and the magnetic field energy density (B 2 /20 ). Joule heating (J 2 ) does not
appear in the MHD system energy density equation because energy lost from the
electromagnetic fields by joule heating [see (??)] increases the internal energy in
the plasma [see (6.29)]; thus, the total MHD energy density, which sums these
energies, remains constant. The terms in the MHD energy flux represent the
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

11

flow of kinetic (m V 2 /2) and internal [P/( 1)] energies with the flow velocity
V, mechanical work done on or by the plasma as it moves (P V), and energy
flow by the electromagnetic fields (EB/0 ) [Poynting vector see (??)].
To illutrate the usefulness of these MHD system conservation equations,
consider the system energy equation (6.36). Integrating this equation over the
volume V of an isolated plasma, the divergence term can be converted using
Gauss theorem (??) into a surface integral that vanishes if there is no flow of
plasma or electromagnetic energy across the surface that bounds the volume.
For such an isolated system the integral of the system energy over the volume
must be independent of time:
&
'
)
P
B2
m V 2
3
+
+
dx
Wk + Wp = constant,
(6.40)
2
1 20
V
in which
Wk
Wp

m V 2
,
2
V
&
'
)
B2
P
3
+
=
dx
,
1 20
V
=

d3x

plasma kinetic energy,

(6.41)

MHD potential energy.

(6.42)

Thus, in the MHD model while there can be exchanges of energy between the
plasma kinetic, and internal and magnetic energies, their sum must be constant.
For a plasma motion to grow monotonically (as in a collective instability), increases in plasma kinetic energy due to dynamical motion of the plasma must
be balanced by reductions in the potential (plasma internal plus magnetic field)
energy in the plasma volume. In Chapter 21 the constancy of the total system
energy in MHD will be used as the basis for developing a variational (energy)
principle for plasma instability, which can occur for a plasma perturbation that
reduces the system potential energy Wp .

6.3

MHD Equilibrium

In this section we discuss the equilibrium (/t = 0) consequences of the system


conservation relations for MHD (6.34)(6.36). In equilibrium the mass density
equation yields m V = 0. In one dimension (x), this equilibrium continuity equation yields m (x)Vx (x) = constant. Thus, in a one-dimensional flow
situation the mass density will be higher (lower) where the flow velocity V is
lower (higher). Equilibrium flows are negligible in MHD for many plasma situations; then the equilibrium continuity equation is trivially satisfied for any mass
density profile m (x).
Next, consider the stress-induced forces which contribute to the system momentum conservation equation (6.35). Consider first the magnetic (subscript
B) contribution that is represented by the JB force density in the momentum equation (6.27) and the magnetic field part of the system stress tensor T
in (6.37). The stress in the magnetic field exerts a force density fB on a fluid
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

12

Figure 6.1: Schematic illustration of the stresses and force densities on a fluid
element of plasma in the MHD model: a) isotropic expansive pressure stress
TP = P I, b) anisotropic magnetic stresses TB , c) pressure gradient force density
N curvature)
fP = P , and d) magnetic force density fB in the normal (
B) directions.
and binormal (
element (infinitesimal volume of MHD plasma fluid) given by
fB

1
B

(B)B = b(B
b)
0
0
2
B
B b(

b)
b)
= b(B
0
0
,
& 2'
b

B
B2
B2
b
TB , (6.43)
+
=
=
I
20
0
20
2
JB =

in which we have used vector identities (??), (??), (??), (??), (??), (??) and
(??). The corresponding force density fP due to the plasma pressure is
fP P = P I TP .

(6.44)

These stresses and force densities are illustrated schematically in Fig. 6.1 and
discussed in the next few paragraphs.
B/B as the base vectors for
ey , b
Consider first the stresses. Adopting
ex ,
a local magnetic field coordinate system, the sum of the pressure and magnetic
stress tensors can be written (in matrix notation) as

ex
0
0
.
/ P + B 2 /20


ey
0
0
P + B 2 /20
ex
ey b
T P + TB

0
0
P B 2 /20
b
"" b,

ex +
ey T
ey + bT
(6.45)
=
ex T
with

T P + B 2 /20 ,

T"" P B 2 /20 .

(For simplicity of presentation, often the directional vectors are omitted and
only the elements of the matrix of tensor coefficients are shown.) The plasma
pressure produces an isotropic tensor (I) expansive (positive) stress, which represents the thermal motion of particles expanding uniformly in all directions. The
magnetic stress is anisotropic. From TB and (6.45), we see that the magnetic
ey perpendicular to the magnetic
stress is expansive (positive) in directions
ex ,
but in tension (negative) along magnetic field lines. Physically,
field B = B b,
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

13

the magnetic field can be thought of as providing a magnetic pressure B 2 /20


perpendicular to magnetic field lines, and tension along field lines as if the
magnetic field lines are elastic cords with tension stress of B 2 /0 along B pressing against the plasma fluid, which is trying to expand perpendicular to the
magnetic field lines due to the combination of the pressure and magnetic energy
density expansive forces.
The force density on an MHD fluid element is given (for subsonic flows where
the Reynolds stress tensor m VV is negligible) by the divergence of this stress
tensor:
f P + fB

P + JB = (TP + TB )
& 2'
(B )B
B
+
= P
20
0
& 2'
B
B2
= P
.
+
20
0

(6.46)

In the last form, the P term represents the isotropic, pressure gradient force,
the next term represents the perpendicular (to B) force due to the magnetic
pressure B 2 /20 and the last term represents the force due to the parallel
tension of magnetic field lines, as if each magnetic cord presses on the fluid
2
with a force density of (B 2 /0 ) = (B 2 /0 )RC /RC
where RC is the local
radius of curvature vector [see (??)] of a magnetic field line.
An MHD fluid element will be in force balance equilibrium, which is usually
just called equilibrium in MHD, if the force density fP + fB vanishes. Then,
there is no net force to drive an inertial force response via the MHD momentum
equation (6.27) and the system momentum conservation equation (6.35) is satified in equilibrium [(m V)/t = 0]. When there is no gradient in the plasma
pressure (an unconfined plasma), the force balance equilibrium becomes
fB = JB = 0,

force-free equilibrium with P = 0.

(6.47)

In order for a magnetic field system to be able to support a pressure gradient in


force balance equilibrium, the current and magnetic field must not be parallel
to each other; rather, their cross product must satisfy
JB = P,

MHD force-balance equilibrium.

(6.48)

Taking the cross product of B with this equation, we obtain the diamagnetic
current J = (BP )/B 2 in (6.8), which is the sum of the diamagnetic flows
of all species of charged particles in the plasma given in (??). The perpendicular
b
)] component of the MHD force-balance equation can also be written
[b(
[from the last form of (6.46)] as
= ln B +

0
P,
B2

perpendicular equilibrium in MHD.

(6.49)

This formula is the same as (??) given previously in Chapter 3 for the magnetic
field curvature if we use the MHD equilibrium condition JB = P .
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

14

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

Because the force density on the plasma is different in different directions, it


is of interest to explore its forms and implications in various relevant directions.
Since the magnetic field direction and curvature are two obviously important
directions, a convenient coordinate system is the Frenet coordinate system whose
orthogonal base vectors for a vector field (B here) are (see Section D.6)
B/B,

Tb

N /,

B
T
N = b/,

(6.50)

T), normal (
N, or curvature) and binormal
which are unit vectors in the tangent (
B) directions of the B field. Decomposing the MHD force density on a fluid
(
element into its components in these orthogonal directions, we find
(
&
%
'
B2
B2
B2

N )(P +
)
B (
B ) P +
.
fP + fB = b (b P ) N (
20
0
20
(6.51)
The conditions for MHD force-balance equilibrium are thus (see Fig. 6.1)

along B:
curvature direction:
binormal direction:

P = P ,
0=b
) '
&
B2
B2
0=
N P +

,
20
0
&
'
B2
B P +
.
0=
20

(6.52)
(6.53)
(6.54)

Since there is no magnetic force along the magnetic field (BfB = BJB =
0), in order to satisfy the first (parallel) MHD force balance condition the plasma
pressure P must be constant along magnetic field lines. (The axial confinement
of plasma in a magnetic mirror is achieved via anisotropic pressure see Problem 6.11.) When nested magnetic flux surfaces exist (see end of Section 3.2),
P/) = 0 requires that the pressure be a function only of the magnetic flux :
P = P ()

B P = (B )

dP
= 0,
d

(6.55)

which vanishes (assuming finite dP/d), by virtue of the condition for the existence of a magnetic flux function (??): B = 0. Further, from the dot
product of the current J with the MHD equilibrium force-balance condition
(6.48) we find
J P = 0.
(6.56)

From these last two equations we see that the vector fields J and B both lie
within, and do not penetrate, magnetic flux surfaces. Further, we see from
(6.48) that in force balance equilibrium the cross product of these two vectors
in the flux surface must equal the pressure gradient, which is perpendicular to
the flux surface (see Fig. 6.2):
JB = P () =
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

dP
.
d

(6.57)

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

15

Figure 6.2: In ideal MHD equilibrium the cross product of the current density J
and magnetic field B vectors within a flux surface is equal to P = (dP/d),
which is normal to the flux surface.

Figure 6.3: Pressure P and magnetic energy density B 2 /20 profiles for: a)
<< 1, and b) $ 1.
When there is no magnetic field curvature, the force balance equilibrium
condition is the same in all directions perpendicular to the magnetic field:
&
'
B2
= 0, MHD equilibrium with no B field curvature. (6.58)
P +
20
To illustrate the implications of this equation, consider the MHD equilibrium of
= B0
ez , For a
a localized plasma placed in a uniform magnetic field B = B0 b
given plasma pressure profile P (x ) that varies in directions (x ) perpendicular
to the magnetic field but does not extend to infinite dimensions, (6.58) yields
& 2'
4
B
P

= B(x ) = B0 1 (x ) .
(6.59)
=
x 20
x
Here, we have defined the very important MHD parameter by
(
. n /%
P (x )
ni
e
= 4.0 1025
(eV)
+
T
(eV)
,
T
(x ) 2
e
i
B0 /20
B2
ne
ratio of plasma pressure to magnetic energy density.

(6.60)

Thus, in an MHD equilibrium, for a situation where the magnetic field B


has no curvature, the plasma digs a magnetic well (region of reduced magnetic
energy density) that is just deep enough so that the sum of the plasma pressure P and magnetic field energy density B 2 /20 is constant (at B02 /20 ) in
all directions perpendicular to the magnetic field. This result is illustrated in
Fig. 6.3 for a cylindrical plasma where the plasma pressure vanishes at r = a for
two cases: small and near unity . The cylindrical form of (6.59) can also be
obtained directly (see Problem 6.14) from the radial force balance equation by
calculating the radial variation of the magnetic field Bz (r) using the azimuthal
component of Amperes law.
When the magnetic field has curvature, the force balance condition in the
normal (curvature) direction is changed to condition (6.53). Then, the pressure
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

16

gradient in the curvature direction can be supported in force balance equilibrium


by either the curvature-induced force density (B 2 /0 ) or the gradient in the
magnetic energy density, or by some combination thereof. When the plasma
pressure is low ( << 1), the magnetic field curvature is equal to the gradient
of the magnetic field energy density [the situation for a vacuum magnetic field
see (??)] plus a small correction due to the plasma pressure. In the limit where
the magnetic field curvature is weak (radius of curvature RC much greater than
the presssure gradient scale length LP P/| P |), the curvature effects are
small and the variation in magnetic field strength is still approximately as given
in (6.59). [In an axisymmetric tokamak both of these small corrections to (6.59)
are unfortunately comparable in magnitude see Chapter 20.] In the binormal
direction, (6.54) shows that in force balance equilibrium, even with curvature
in the magnetic field B, P + B 2 /20 is constant in the binormal direction
increases in the plasma pressure P in the binormal direction are balanced by
decreases in magnetic energy density B 2 /20 , like in (6.59).
From the preceding discussion is is clear that the parameter characterizes
the relative importance of the plasma pressure P versus the magnetic field B.
For << 1 the plasma pressure has a small effect on the MHD equilibrium
and the magnetic field structure is approximately that determined from a vacuum magnetic field representation (??). Also, the diamagnetic current is small
(J $ B/20 ), as is the (diamagnetic) magnetic susceptibility due to the
plasma magnetization produced by the magnetic moments of all the charged
particles in the plasma gyrating in the magnetic field [M $ /2 see discussion after (6.12)]. Since the magnetic field is much stronger than the plasma
pressure in this regime, it can be used to provide a magnetic bottle for plasma
confinement. In the opposite limit ( >> 1) where the the plasma pressure in
much larger than the magnetic energy density, in general the plasma pushes the
magnetic field around and carries it along with its natural motions (pressure
expansion plus flows). A key question for magnetic fusion confinement systems
is the maximum they can stably confine in equilibrium; the 510% that is
needed for economically viable deuterium-tritium fusion reactors is apparently
accessible in many types of toroidal confinement systems see Chapter 21.
It is often asked: can a finite pressure plasma support itself entirely with the
diamagnetic current and the magnetic field it produces, without any externally
imposed magnetic field? That is, can a plasma organize itself into a closed
magnetic equilibrium that has no connection to the outside world? In order
to examine this question, we consider the equilibrium [(m V)/t = 0] MHD
sytem momentum (or force balance) equation obtained from (6.35): T = 0.
Taking the dot product of this equation with the position vector x from the
centroid of the plasma system (to obtain a measure of the MHD system potential
energy density), we obtain the relation
0 = x T = (x T) x : T = (x T) tr{T},

(6.61)

in which we have used vector identities (??), (??) and (??). Integrating this

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

17

last form over a volume larger than the proposed isolated plasma, we obtain
)
))
! dS (x T) = d3x tr{T},
(6.62)
S

in which we have used the tensor form of Gauss divergence theorem (??) to
convert the volume integral to a surface integral. We now examine the integrals
on the left and right separately. For the integral on the left we assume negligible
flows (V 0) and use (6.37) for T. Then, the integral on the left can be written
as
'
(
% &
))
))
(x B)B
1
B2
r
! dS (x T) = ! dS x P +
= 0.

3
20
0
r
(6.63)
As indicated at the end, in the limit of large radial distances r from the isolated
plasma this integral vaishes because since there are apparently no magnetic
monopoles in the universe, the magnetic field B must decrease like that for a
dipole field does (|B| 1/r3 ) so the integrand scales as 1/r5 and when integrated
over the surface (|dS| 4r2 ) one finds that the integral decreases at least as
fast as 1/r3 . Next, we consider the integral on the right. Using the matrix
definition of the stress tensor T given in (6.45), we find (for an isolated plasma
within a finite volume V )
&
'
)
)
B2
3
3
d x tr{T} =
d x 3P +
= constant.
(6.64)
20
V
V
The only way this last integral can vanish, as is required by the combination
of (6.62) and (6.63), is if the plasma pressure P and magnetic energy density B 2 /20 (both of which are intrinsically positive quantities) vanish. Thus,
we have found a contradiction: no isolated finite-pressure plasma can by itself develop a self-confining magnetic field in force balance equilibrium. This
proof
is sometimes called a virial theorem (because it results from
5
5 3 and analysis
d x x f = d3x x T = 0) and was first derived by V.D. Shafranov.1

6.4

Boundary Conditions and Shock Relations

The basic subject to be discussed here are the jump conditions at a discontinuity in a plasma or at a plasma-vacuum interface, and then the corresponding
bounday conditions at a vacuum wall or around coils for a free-boundary equilibrium. See Section 3.2 of the Freidberg book. These same equations become
the shock conditions in a plasma. This section will be written later.

6.5

MHD Dynamics

To explore the elementary dynamical (evolution in time) properties of a plasma


in the MHD model, we first assume that the plasma fluid moves with a velocity
1 V.D. Shafranov, in Reviews of Plasma Physics, edited by M.A. Leontovich (Consultants
Bureau, New York, 1966), Vol. II.

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

18

V(x, t) and determine the changes in the mass density m , pressure P and magnetic field B induced by V. Then, these responses are used in the momentum
equation (6.27) which is then solved self-consistently to determine the mass flow
velocity V.
We begin by considering the temporal evolution of the mass density in response to V, which is governed by (6.25):
m /t|x = Vm + m V

dm /dt = m V,

(6.65)

in which we have used the vector identity (??) in obtaining the first form and the
total time derivative definition in (6.30) in obtaining the second form. Here, as
shown in Fig. ??, in the Eulerian (fixed position) picture [first form of (6.65)], the
flow causes changes in the mass density at a fixed point by advecting (Vm )
the mass flow at velocity V into a region of different mass density, or by compressibility (V ,= 0) of the flow. In the Lagrangian (moving with fluid element) picture [second form of (6.65)], the mass density only changes due to the
compressibility of the flow (V ,= 0).
The pressure evolution can be determined from the isentropic form of the
MHD equation of state [i.e., (6.29) neglecting the small entropy production due
to joule heating]:
P
1 dP
dm
1 dP
d
ln
=

=
+ V = 0,
dt m
P dt
m dt
P dt

(6.66)

in which (6.65) has been used to obtain the last form. With the total time
derivative definition (6.30), this yields

in which

P
= VP P V = VP c2S m V
t
cS

4
P/m ,

MHD sound speed (m/s).

(6.67)
(6.68)

Thus, like the mass density, the plasma pressure changes in MHD are due to
advection (VP ) and flow compression (V ,= 0). The presence of the sound
speed in the last form of (6.67) shows that the compressiblity of the flow leads to
pressure changes that move at the MHD sound speed through the plasma. Thus,
the fluid motion at velocity V causes advection and compressibility changes in
the mass density m and plasma pressure P , which are scalar quantities.
Note that the MHD sound speed is different from the ion acoustic speed
(??) in Section 1.4 because in a MHD description both the electrons and ions
have fluidlike (inertial) responses whereas for ion acoustic waves while the ions
have a fluidlike response the electrons respond adiabatically. Unfortunately,in
plasma physics the same symbol is usually used for both wave speeds which
is meant is usually clear from the context. Also note that for most plasmas
with comparable electron and ion temperatures these two speeds are close in
magnitude.
The next question is: what is the effect of the fluid motion on the magnetic
field B(x, t), which is a vector field? Physically, we know that plasmas have
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

19

a very high electrical conductivity (low resistivity). In the ideal MHD model
we set the resistivity to zero and hence effectively assume infinite electrical
conductivity; thus, the plasma is a superconductor in ideal MHD. From the
properties of a superconducting wire of finite cross-section, we know that the
magnetic field is frozen into it and moves with the wire as it is moved. Thus,
we can intuitively anticipate that a fluid element in our superconducting ideal
MHD plasma will carry the magnetic field (or at least the bundle of magnetic
field lines penetrating it) with it wherever it moves and will always contain the
same amount of magnetic flux (number of field lines2 ). We can also anticipate
that the addition of resistivity in the resistive MHD model will allow some
slippage of the magnetic field lines relative to the fluid element.
We now develop mathematical representations of the idea that the magnetic
field is mostly frozen into an MHD fluid element
55 and moves with it. Consider
the time derivative of the magnetic flux S B dS [see (??)] though an
open surface S in the fluid that moves with the fluid at velocity V:
(
))
)) %
d
d
d
dB
=
dS + B (dS) .
(6.69)
B dS =
dt
dt S
dt
dt
S
The total time derivative is appropriate here because we are seeking the change
in the magnetic flux penetrating a (changing) surface whose boundary is distorted in time as it moves with the fluid velocity V(x, t), which is in general
nonuniform. The time derivative of the (vectorial) differential surface area dS
represents changes due to changes in its constituent differential line elements
induced by the nonuniform flow see Section D.4. Using (??) for this time
derivative and the definition of the total time derivative in (6.30), we find
'
(
)) %&
B
d
=
+ VB + B (V) B V dS
dt
t
S
(
)) %
B
(VB) dS,
(6.70)
=
t
S
in which we have used vector identity (??) and the Maxwell equation B = 0
in going from the first to the second line.
For the evolution of the magnetic field B we use Faradays law (6.31) together
with the MHD Ohms law (6.28) to specify the electric field E:
2
B
= E = (VB) J $ (VB) +
B
t
0
MHD magnetic field evolution.
(6.71)
Here, in the last, approximate form we have used J = B/0 (Amperes law),
neglected for simplicity, and used the vector identity (??) and the Maxwell
2 While magnetic field lines do not really exist since their properties cannot be measured,
they are a useful concept for visualizing the behavior of the magnetic field B.

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

20

equation B = 0. Substituting this magnetic field evolution into (6.70), using


Amperes law for J again and Stokes theorem (??), we finally obtain
))
6

d
=
dS J =
d# B.
(6.72)
dt

0
S
C
In ideal MHD where 0, this becomes
d
= 0,
dt

ideal MHD frozen flux theorem.3

(6.73)

Thus, in the absence of resistivity the magnetic flux (number of field lines)
through an open surface that moves with the fluid velocity V is frozen into
the fluid and hence constant: the magnetic field moves with the superconducting
ideal MHD fluid just as we wanted to prove! The key ingredient in this derivation
is the VB term in the MHD Ohms law. It led to the (VB) term in
the magnetic field evolution equation (6.71) and causes the magnetic field to be
carried along with the ideal MHD fluid. Hence, this (VB) term represents
the advection of the vector field B by the flow velocity V; note that this vector
field advection operator is different in structure from the advection operator
for scalar quantities such as the mass density (Vm ). Since the MHD
Ohms law is an approximation to the electron momentum balance equation, it
is fundamentally the electron fluid into which magnetic field is frozen (despite
the fact that the advection is induced by the overall plasma mass flow velocity
V).
By taking the limit of an infintesimally small surface S in the preceding
derivation, one can show that an individual magnetic field line is carried along
with the superconducting ideal MHD plasma. This can also be shown directly
by examining the conditions under which the time derivative of the definitions
of magnetic field lines vanish see Problems 6.15 and 6.16. However, it is
important to note that all these derivations have some ambiguity because the
labeling of a magnetic field line is not unique [see discussion after (??)] and
the properties of magnetic field lines cannot be measured. Thus, while we can
mark infintesimal elements of a fluid (e.g., with radioactive nuclei or fluorescing
partially ionized atoms), and know that the magnetic field is frozen into the ideal
MHD fluid elements as they move, the association with a particular magnetic
field line from one instant in time to the next is not unique. The frozen flux
methodology provides a prescription for labeling field lines as they move. While
it is not a unique prescription, it represents a very important tool for visualizing
the motion of magnetic fields in a moving plasma in the MHD model.
The frozen flux theorem provides a very strong constraint on the motions
of the magnetic field in an ideal MHD plasma. In particular, in this model
adjacent magnetic field lines and flux bundles that are originally adjacent to
each other will forever remain adjacent. Also, magnetic flux bundles and fluid
3 This theorem is also known as the Alfv
en frozen flux theorem. It is the magnetic field
analogue of the Kelvin circulation theorem (??) for the constancy of the circulation or vorticity
flux in a vortex in an inviscid neutral fluid.

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

21

Figure 6.4: Possible MHD evolution of a set of field lines in a sheared slab
magnetic field model: a) initial sheared magnetic field equilibrium, b) sinusoidal
perturbation in ideal MHD ( = 0), and c) resistive MHD ( ,= 0) with magnetic
field reconnection into magnetic island structures.
elements are tied together, cannot break up or tear, and cannot interchange
positions relative to each other. Thus, as illustrated in Figure 6.4, in the ideal
MHD model the topology of magnetic field lines and flux surfaces is conserved
nested magnetic flux surfaces remain forever nested (even though their shape
may become highly distorted), and plasma in regions inside (or outside) a
given magnetic flux surface remain inside (outside) forever. The inclusion of
resistivity in the MHD model allows diffusion of the magnetic field relative to
the plasma, and hence reconnection of the magnetic field lines and changes in
the magnetic topology for example by forming a magnetic island such as
indicated in Figure 6.4c. In section 6.7 we discuss the relative importance of
resistivity in MHD analyses of plasmas.
The most convenient form of the MHD momentum equation (6.27) for dynamical analyses uses the middle form of the force density fB in (6.46) and is
given by
&
'
B2
dV
(B )B
= P +
+
.
(6.74)
m
dt
20
0
Note that we have now reduced the full MHD equation set (6.25)(6.44) to
just three (or seven component) equations the scalar pressure equation in
(6.67), the vector magnetic field evolution equation in (6.71) and this last vector
momentum equation (6.76). These equations are usually all we need to describe
the linear and nonlinear dynamics of plasmas in the MHD model. [The mass
density equation (6.65) is only needed when the equilibrium mass density is
inhomogeneneous.] Note that for these MHD dynamical model equations the
charge continuity equation J = 0 is automatically satisfied by our having
used Amperes law to replace the current J with B/0 , which is divergence
free. Also, the electric field E does not appear because it was replaced by
VB + J using the MHD Ohms law.

6.6

Alfv
en Waves

To illustrate the fundamental wave responses of plasmas in the MHD model


(Alfven waves named after their discoverer), we consider plasma responses
to small perturbations in the simplest possible plasma and magnetic field model.
Namely, for the equilibrium we consider a uniform, nonflowing (V0 = 0) plasma
This model
ez = B0 b.
in an infinite, homogeneous magnetic field B0 = B0
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

22

trivially satisfies the MHD equilibrium force balance condition (6.48) since
0 J0 = B0 = 0 and P = 0 because both the equilibrium magnetic field
B0 and pressure P0 are uniform in space. For perturbed responses we assume
m = m0 + m ,

P = P0 + P ,

V = V,

B = B0 + B,

(6.75)

in which the zero subscript indicates equilibrium quantities and the tilde over
quantities indicates perturbed variables. Decomposing the perturbed mag = B

" = b(
" b]
b
B)
and perpendicular [B
netic field into its parallel [B
components, we find the square of the magnetic field strength B
b
B)]
b(
is
(B0 + B)
= B 2 + 2B0 B
|2 $ B 2 + 2B0 B
" + B
2 + |B
" . (6.76)
B 2 (B0 + B)
0
0
"
We will use the last expression, which is the linearized form (i.e., it neglects
terms that are second order in the perturbation amplitudes).
Substituting the equilibrium plus perturbed quantities in (6.75) and (6.76)
into the ideal MHD equations for the evolution of the pressure (6.67), flow
velocity (6.74) and magnetic field [(6.71) with 0] and linearizing (neglect
second and higher order terms in the perturbation amplitudes), we obtain
P
t

V
m0
t

B
t

= P0 V,
,
"
B
B
1
0

= P +
(B0 )B,
+
0
0

= (VB
0 ) = B0 ( V) + (B0 )V.

(6.77)
(6.78)
(6.79)

In the last equation we used vector identity (??) and set to zero terms involving
gradients of the homogeneous equilibrium magnetic field B0 . Equations for the
parallel and perpendicular components of the magnetic field are obtained from
the corresponding projections of the magnetic field evolution equation:
+ (B0 )V" ,
" /t = B0 (V)
B
.
/t = (B0 )V
B

(6.80)
(6.81)

These equations can be combined into a single (vector) equation by taking the
partial derivative of the perturbed momentum equation (6.78) and substituting
in the needed partial derivatives from the other equations (see Problem 6.20):

2V
+ c2 [2 V

= (c2S + c2A )(V)


A
" " V" b" ( V)]
2
t

(6.82)

in which
cA

B0
B0
$ 2.2 1016
m/s,
0 m0
ni Ai

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Alfven speed.

(6.83)

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

23

Here, Ai mi /mp is the atomic mass value of the ions, the perpendicular ()
and parallel (0) subscripts indicate the respective components of the quantities
as defined in (??)(??). The magnitude of the Alfven speed can be appreciated
by noting its relationship to the sound speed defined in (6.68):
P0 /m0

c2S
= 2
= .
c2A
B /0 m0
2

(6.84)

Thus, for < 1 the Alfven speed is a factor of about 1/ greater than the
MHD sound speed.
While (6.82) clearly has a wavelike structure, it is a quite complicated and
anisotropic wave equation. We consider here only some special cases to illustrate
the basic waves involved. (Section 7.6* provides a comprehensive analysis.)
First, consider waves propagating purely perpendicular to the magnetic field
by setting " = 0. Then, taking the divergence of (6.82) we obtain
% 2
(

2
2
2
2

(c
+
c
)

= 2 = k
(c2A + c2S ),
A
S
( V ) = 0
t2
compressional Alfven waves.
(6.85)
This wave equation describes fast compressional Alfven waves. In the last
exp[i(kxt)] to obtain the wave
form we assumed a wave-like response V
dispersion relation. Compressional Alfven waves propagate perpendicular
to
4
the magnetic field with a wave phase speed given by V /k = c2A + c2S ,
which is the fastest MHD wave phase speed. These waves propagate by per ,= 0) and also involve magnetic field
pendicular flow compression ( V

compression [B" ,= 0 see (6.80)] and pressure perturbations [P ,= 0 see


(6.85)]. Adding the pressure perturbation (6.77) and B0 /0 times the magnetic
perturbation (6.80) with " = 0, one can show that
,
,
"
"

B
B
B
B

V
2
0
0
2
2
2
2
2
= (cA + cS ) P +
P +
= (cA + cS ) m0
t2
0
t
0
(6.86)
in which for the last form we have used (6.78) with " = 0. Thus, the compressibility in the perpendicular flow also causes the sum of the perturbed pressure
and magnetic field energy density to satisfy a compressional Alfven wave equation. Physically, as can be noted from the importance of the perpendicular
component of (6.78) in these waves, the compressional Alfven waves are the
responses of the plasma to imbalances in the perpendicular (to B) force balance
in the plasma. Thus, on equilibrium time scales (after these wave responses
have propagated away), MHD plasma responses will be in radial force balance
equilibrium and not have any driving sources for compressional Alfven waves:
J0 B0 = P0 ,

= 0,
V

" /0 = 0.
P + B0 B

(6.87)

These are the lowest order conditions for equilibria and perturbations in an
MHD plasma (even in inhomogeneous magnetic fields see Chapter 21); they
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

24

V,
P ) in the three fundamental types of MHD
Figure 6.5: Perturbations (B,
waves: a) compressional Alfven, b) shear Alfven, and c) sound.
obtain for time scales
4 long compared to the fast compressional Alfven wave
period: t >> 1/k c2A + c2S .
= 0) MHD waves propagating purely
Next, consider incompressible (V
along the magnetic field ( = 0). Then, the perpendicular component of the
general MHD wave equation (6.82) becomes
'
& 2

2
2

= 2 = k"2 c2A , shear Alfven waves. (6.88)


A " V = 0
t2
These are called slow Alfven waves because their (parallel) phase speed V"
/k" = cA is less than the phase speed for the compressional Alfven waves.
induces a
They are called shear (or torsional) Alfven waves because their V

perpendicular magnetic field perturbation B that shears or twists the magnetic


field see (6.81). In the MHD model, instabilities often arise that indirectly
excite shear Alfven waves; such instabilities must have exponential growth rates
Im{} > k" cA so they are not be stabilized by the energy required to excite
these shear Alfven waves.
= V" b)
propaFinally, consider compressible waves in the parallel flow (V
gating along the magnetic field ( = 0). Then, the parallel component of the
general MHD wave equation (6.82) becomes
'
& 2

2 2
cS " V" = 0 = 2 = k"2 c2S , parallel sound waves. (6.89)
t2
These are neutral-fluid-type sound waves (see A.6) that propagate along the
magnetic field by parallel compression of the flow (" V" ,= 0). They are electrostatic waves since, as can be seen from (6.80) and (6.81), they produce no
= 0 for these waves). MHD instabilities often
magnetic perturbations (i.e., B
indirectly excite parallel sound waves; such instabilities must have exponential
growth rates Im{} > k" cS so they are not be stabilized by the energy required
to excite the sound waves.
The properties of the perturbations in these three fundamental types of
MHD waves are illustrated in Fig. 6.5. As shown in Fig. 6.5a, (fast) compressional Alfven waves have: oscillatory parallel magnetic field perturbations
" that increase or decrease the local magnetic field strength (density of field
B
lines), compressible pependicular flows, and corresponding oscillatory pressure
perturbations, all in the direction perpendicular to the equilibrium magnetic
ez , which is horizontal in the figure. In contrast, the
field direction B0 = B0
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

25

(slow) shear Alfven waves (Fig. 6.5b) have: oscillatory perpendicular magnetic
along the magnetic field, but
and oscillatory perpendicular flows V
fields B
no pressure perturbation (because these perurbed flows are incompressible). Finally, as shown in Fig. 6.5c, the parallel sound waves have: no magnetic field
perturbation (because they are electrostatic), an oscillatory compressible parallel flow V" and corresponding pressure P perturbations along the magnetic field
direction.
In the more general case of propagation of MHD waves at arbitrary angles
to the magnetic field direction, these three types of waves become coupled (see
Section 7.6). These waves also become coupled in inhomogeneous magnetic fields
because the parallel and perpendicular directions vary spatially. Nonetheless,
the basic wave characteristics we have discussed are usually still evident in these
more complicated situations.

6.7

Magnetic Field Diffusion in MHD

In order to examine the effect of electrical resistivity on a plasma in the MHD


model, consider first the evolution of the magnetic field in (6.71) without the
advection term:
2
B
=
B,
t
0

magnetic field diffusion equation.

(6.90)

This equation describes the diffusion (see Section A.5) of the magnetic field
(both its magnitude and directional components) that is caused by the electrical
resistivity of a plasma. The diffusion coefficient is
&
'&
'

me e
Zi
ln
3
=
$ 1.4 10
m2 /s
D =
0
0 ne e2
17
Te (eV)]3/2
magnetic field diffusivity.
(6.91)
Phenomenologically, since we can write D = e (c/pe )2 , magnetic field diffusion can be thought of [via D (x)2 /t see (??)] as emanating from a
random walk process in which magnetic field lines step a collisionless skin depth
(x c/pe ) in an electron collision time (t 1/e ). The relative magnitude
of the magnetic field diffusivity can be ascertained from its relationship to the
classical diffusivity D defined in (??):
&
'
e #2e

D
Te + Ti
ne (Te + Ti )
(6.92)
=
= .
=
2
2
2
/0
e (c/pe )
2Te
c (0 B
2
Thus, for a plasma with < 1 particles diffuse classically across magnetic
field lines slower than the magnetic field lines themselves diffuse relative to
the plasma! However, in most plasmas of interest microscopic turbulence in
plasmas causes an anomalous perpendicular transport that is rapid compared
to the magnetic field diffusion; hence one can usually consider the magnetic field
to be stationary for calculations of anomalous transport.
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

26

To illustrate the spatial and temporal scale lengths involved in magnetic diffision, consider the distance an electromagnetic wave can penetrate (see Section
1.5) into a resistive medium in which the magnetic field behavior is governed by
exp[i(kx t)], the diffusion equation
(6.90). For wavelike perturbations B
becomes
4
4
= k = i0 / = (1 + i) (/2)(0 /). (6.93)
= k 2 (/0 )B
i B

To use the analysis of Section 1.5, we identify this complex wavenumber k as


the transmitted wavenumber kT in (??). Thus, an electromagnetic wave will be
dissipated and damped exponentially, as it oscillates spatially (due to Re{kT })
and propagates into a resistive medium, with a characteristic decay length of
7
1
2
=
, resistive skin depth.
(6.94)

Im{(kT }
0

It is called a skin depth because of its analogy with the problem of determining how far an oscillating magnetic field (e.g., due to 60 Hz AC electricity)
penetrates into a cylindrical wire of finite radius. This skin depth formula is
appropriate for radian frequencies < e , while the collisionless skin depth formula (??) is appropriate for higher frequencies see Problem 6.25. For Te =
2000 eV, which gives /0 $ 0.016 m2 /s (close to the resistivity of copper at
room temperature of /0 $ 0.135 m2 /s), the resistive skin depth ranges from
0.07 mm for f = /2 = 104 Hz ( = 2 104 ) to about 1 cm for 60 Hz.
Another way of illustrating the temporal behavior of magnetic field diffusion
in a magnetized plasma is to ask: on what time scale will a magnetic field
component diffuse away from being localized to a region of width L ? Because
for diffusive processes the diffusion coefficient scales with spatial and temporal
steps as D (x)2 /t L2 / (see Appendix A.5), we can estimate phenomenologically that L2 /(/0 ). One often considers a cylindrical model
consisting of a column of magnetized plasma with radius a that initially carries
an axial current. For such a cylindrical model the resistivity-induced decay time
of the current (and induced azimuthal magnetic field) is (see Section A.5)
$

a2
,
6 /0

resistive skin diffusion time.

(6.95)

Here, the numerical factor of 6 is a cylindrical geometry factor which more pre2
$ 2.4052 $ 5.78
cisely is the square of the first zero of the J0 Bessel function: j0,0
see Appendix A.5 and (??). However, the additional accuracy is unwarranted
both because of the approximations involved in the simple model used to derive
and because of the intrinsic accuracy of the electrical resistivity ($ 1/ ln
510%). For a plasma of radius a = 0.3 m with Te = 2000 eV, which gives
/0 $ 0.016 m2 /s, the skin time is 1 s.
Finally, we discuss the relative importance of the two contributions to magnetic field evolution (6.71) in the MHD model: advection of the magnetic field
by (VB), and resistive diffusion by (/0 )2 B. The relative importance
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD


of these two terms is indicated by the scaling properties of their ratio:
&
'
2
3/2
cA /L"
17
|(VB)|
13 a B[Te (eV)]

,
$ 1.6 10
S=
|(/0 )2 B|
(/0 )/a2
ln
L" Zi ni Ai
Lundquist number.4

27

(6.96)

Here, we have taken the typical velocity to be the Alfven speed cA and assumed
scale lengths L" [e.g., periodicity scale length along B see (6.81)] for the
advection process and a (e.g., plasma radius) for the magnetic diffusion. Typical
Lundquist numbers range from 102 for cold, resistive plasmas, to 105 1010 for
the earths magnetosphere and magnetic fusion experiments, to 1010 1014 for
the suns corona and astrophysical plasmas.
Because the Lundquist number is large for almost all magnetized plasmas
of interest (and extremely large for high temperature plasmas), one might be
tempted to just set the resistivity to zero (S ) and always use the ideal
MHD model. Indeed, throughout most of a plasma the magnetic field is frozen
into and moves with the plasma fluid. However, a small resistivity can be
very important in resistive boundary layers. The boundary layers occur in
in
the vicinity of magnetic field lines where the parallel derivative (B0 )V
evolution becomes dominated by resistive evolution
(6.81) vanishes so the B
, rather than by advection. The width of these resistive boundary layers
of B
scales inversely with a fractional power of the Lundquist number (S 1/3 or
S 2/5 ) and hence is not negligible see Chapter 22. Since resistivity allows the
magnetic field lines to slip relative to the plasma fluid, they relax (in the resistive
layers) the frozen flux constraint and thereby allow new types of instabilities
resistive MHD instabilities, which are described in Chapter 22. Since the
resistivity only relaxes the frozen flux constraint in thin layers, resistive MHD
instabilities grow much slower (by factors of S 1/3 or S 3/5 ) than ideal MHD
instabilities. However, resistive MHD instabilities are quite important, because
they can lead to turbulent plasma transport (see Section 25.3) and because
in these narrow resistive boundary layers the magnetic field lines can tear or
reconnect and thereby lead to changes in the magnetic topology (see Section
22.3). For example, they can nonlinearly evolve into a magnetic island structure
like that shown in Fig. 6.4c.

6.8

Which Plasma Description To Use When?

In this section we discuss which types of plasma descriptions are used for describing various types of plasma processes in magnetized plasmas. This discussion
also serves as an introduction to most of the subjects that will be covered in the
remainder of the book. The basic logic is that the fastest, finest scale processes
4 Many plasma physics textbooks refer to this as the magnetic Reynolds number. However, S is the ratio of linear advection to a dissipative process rather than the ratio of nonlinear
advection to a dissipative process, as the neutral fluid Reynolds number is see (??). We
will call S the Lundquist number to avoid the implication that this dimensionless number is
indicative of nonlinear processes that always lead to turbulence when it is large.

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

28

require kinetic descriptions, but then over longer time and length scales more
fluidlike, macroscopic models become appropriate. Also, the equilibrium of
the faster time scale processes often provide constraint conditions for the longer
time scale, more macroscopic processes.
In a magnetized plasma there are many more relevant parameters, and their
relative magnitudes and consequences can vary from one application to another.
Thus, to provide a table similar to Table ?? for magnetized plasmas, we need to
specify the parameters for a particular application. We will choose parameters
toward the edge (r/a = 0.7) of a typical 1990s large-scale tokamak plasma
(e.g., the Tokamak Fusion Test Reactor: TFTR): Te = Ti = 1 keV, ne = 31019
m3 , B = 4 T, deuterium ions, Zeff = 2, L" = R0 q $ 6 m, a = 0.8 m, Lp = 0.5
m. In a magnetized plasma the unmagnetized phenomena listed in Table ??
still occur; however, their effects only influence responses along the magnetic
field direction. Parameters for the gyromotion, bounce motion and drift motion
of charged particles in this tokamak magnetic field structure are approximately
the same as those indicated in (??) and (??).
Table 6.1 presents an outline of magnetized-plasma-specific plasma phenomena, and their relevant time scales, appropriate models and possible consequences for the tokamak plasma parameters indicated in the preceding paragraph. In it time scales are indicated in half order of magnitudes (100.5 =
3.16 3). As indicated, the fastest magnetic-specific process in magnetized
plasmas is the gyromotion of particles about the magnetic field, for which the appropriate model is the Vlasov equation. The ion gyromotion leads to cyclotron
(Bernstein) waves, finite ion gyroradius (FLR) effects and a perpendicular dielectric response (Sections 7.5, 7.6). There are of course also electron cyclotron
motion and waves. The propagation of (electron and ion) cyclotron-type waves
in plasmas and their use for wave heating of magnetized plasmas are discussed
in Chapters 9 and 10. If the electron or ion distribution function is peaked at a
nonzero energy (so f0 / > 0), it can lead to cyclotron instabilities (Chapter
18) whose nonlinear evolution to a steady state or bursting situation is often
determined by collisions (Section 24.1).
The next fastest time scales are typically those associated with the the Alfven
wave and sound wave frequencies which are described by the ideal MHD model:
(6.25)(6.39) with 0. As indicated in Table 6.1, in the usual situation
where compressional Alfven waves are stable, their effect is to impose radial (
to B) force balance equilbrium [(6.48) and Chapter 20] on the plasma and lower
frequency perturbations in the plasma. The shear Alfven and sound waves
can lead to virulent macroscopic current-driven (kink) and pressure-gradientdriven (interchange) instabilities (Chapter 21). The nonlinear consequences
of an ideal MHD instability is often dramatic movement or catastrophic loss
of the plasma in a few to ten instability growth times; hence most magnetic
confinement systems are designed to provide ideal MHD stability for the plasmas
placed in them.
Next, we turn to the sequentially slower particle and plasma motions along
(0), across () and perpendicular () to the magnetic field B. The fastest
motion along a magnetic field line is the electron bounce motion, which is deDRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

29

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD


Table 6.1: Phenomena, Models For A Magnetized Plasma
Physical Process

Time Scales

cyclotron waves
cyclotron inst.

1/ci 108 s
1/ci 108 s

Alfven waves
compressional
shear
sound waves
parallel (0) to B
electron bounce
electron collisions
ion bounce
Ohms law in MHD
cross () to B
diamagnetic flow
cross flow equil.
perp. () to B
plasma transport
B field evolution

a/cA 107 s
L" /cA 106 s
a/cS 105.5 s

Species,
Plasma Model

Consequences

Vlasov
Vlasov

dielectric resp.
NL, via collisions

ideal MHD
ideal MHD

P = JB
J-driven inst.
P -driven inst.

parallel kinetic
0 Vlasov
drift kinetic
0 Vlasov

ne , Te const. 0 B
"e
, q"e , b
ni , Ti const. 0 B

> 1/e 105 s

resistive MHD

resistive inst.

< 105 s
1/
1/i 103 s

gyrokinetic
drift kinetic

drift wave inst.


cross flow damp.

E a2 /4
a2 0 /6

two-fluid
res./neo. MHD

loss of plasma
B field diffusion,
magnetic islands

1/be 106.5 s
1/e 105 s
1/bi 104.5 s

scribed by a parallel motion version of the Vlasov equation [the drift kinetic
equation (??) without the collision operator and drift velocity vD ]. On time
scales longer than the electron bounce time (1/be ), the lowest order distribution
function becomes constant along field lines (" f0e = 0 and hence density and
temperature become constant along B), and distinctions between trapped and
untrapped electrons and their differing particle orbits become evident. For the
parameters chosen, we have an electron collision length e = vT e /e $ 200 m
33L" and hence e " 33 >> 1. This is a typical toroidal plasma which is
often (confusingly) called collisionless because the collision length is long
compared to the parallel periodicity length. Since the electron gyroradius is
negligibly small, the collisional evolution of the electron species on the collision
time scale (1/e ) is governed by the (electron) drift kinetic equation (??). Its solution for axisymmetric toroidal plasmas is discussed in Section 16.2*. For times
long compared to the electron collision time the plasma acquires its electrical
resistivity and the collisions of untrapped electrons produce entropy through

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

30

"e ) see
neoclassical heat conduction (q"e ) and parallel viscosity (b
Chapter 16. Similarly, the lowest order ion distribution function, density and
temperature become constant along magnetic field lines for time scales longer
than the ion bounce time (1/bi ) and their collisional effects (in relaxing cross
flows within a magnetic flux surface) become evident on the ion collision time
scale (1/i ).
The plasma exhibits an electrical resistivity for time scales longer than the
electron collision time (1/e ). Its introduction into MHD leads to the resistive
MHD model: (6.25)(6.39). Since the introduction of resistivity relaxes the
ideal MHD frozen flux constraint (in narrow layers), it can lead to resistive
MHD instabilities related to their ideal MHD counterparts (kink tearing,
P -driven interchange resistive interchange), which, however, grow more
slowly and hence are less virulent see Chapter 22.
The next set of phenomena concern the effects of particle drifts and plasma
species flows in the cross direction ( perpendicular to B and within a flux
surface if it exists). On this time scale a global (as opposed to local) description
of the magnetic field is usually required. The diamagnetic flows of electrons
and ions lead to drift-wave-type oscillations (Sections 7.4* and 8.6*) and instabilities (Section 23.3*). Since these universal instabilities involve modes with
significant ion gyroradius (FLR) effects (#i k #i 1), the gyrokinetic
equation is used to describe their nonlinear evolution into microsopic plasma
turbulence (Chapter 25) that leads to anomalous radial transport (Chapter 26)
of the plasma. On the same time scales the combination of the EB and diamagnetic flows come into equilibrium (a steady state saturation or bounded
cyclic behavior); flow components within a magnetic flux surface in directions
in which the magnetic field is inhomogeneous (e.g., the poloidal direction in an
axisymmetric tokamak) are damped on the ion collision time scale (1/i ) see
Section 16.3*. Steady-state net radial transport fluxes can only be properly calculated after the flows within magnetic flux surfaces are determined and relaxed
to their equilibium values. Also, in determining transport fluxes it is implicitly
assumed that nested magnetic flux surfaces exist and that radial transport is
to be calculated relative to them.
Finally, we reach the transport time scales on which the plasma and magnetic field diffuse radially out of the plasma confinement region, and radiation
(Chapter 14) can be significant. Plasma transport (relative to the magnetic
field) is usually modeled with two-fluid equations averaged over magnetic flux
surfaces to yield equations that govern the transport of plasmas perpendicular
to magnetic flux surfaces see Chapter 17. However, the radial particle and
heat diffusion coefficients D , are usually assumed to be the sum of those
produced by anomalous transport (Chapter 26) and those due to classical [(??),
(??) and Chapter 15] and neoclassical (Chapter 16) transport processes. For a
cylindrical-type plasma model the characteristic time scale for the usally dominant plasma energy loss is (see Section 17.3) approximately E a2 /4 in
which a is the plasma radius; for the plasma parameters we are considering it
is of order 0.1 s. Simultaneously, the magnetic field is diffusing. The characteristic time scale for diffusive transport of magnetic field lines out of a cylindrical
DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

31

plasma is a2 /(6/0 ); for the plasma parameters we are considering it is of


order 1 s. If resistive or neoclassical MHD tearing-type instabilities are present,
they can reconnect magnetic field lines on rational magnetic flux surfaces and
evolve nonlinearly by forming magnetic islands which grow (to saturation or
total plasma loss) on a fraction ( 0.1) of the magnetic field diffusion time scale
. Since the magnetic field typically diffuses more slowly than energy is lost via
anomalous transport (i.e., >> E or /0 << ), it is usually reasonable
to assume that the magnetic field is stationary and the plasma moves relative
to it via Coulomb-collision-induced or anomalous plasma transport processes.
REFERENCES AND SUGGESTED READING
The MHD description of a plasma and its properties are presented in
Schmidt, Physics of High Temperature Plasmas (1966,1979), Chapts. 3,4 [?]
Krall and Trivelpiece, Principles of Plasma Physics (1973), Chapts. 2.3 [?]
Freidberg, Ideal Magnetoydrodynamics (1987) [?]
Sturrock, Plasma Physics, An Introduction to the Theory of Astrophysical, Geophysical & Laboratory Plasmas (1994), Chapts. 11,12 [?]
Hazeltine and Waelbroeck, The Framework of Plasma Physics (1998), Chapts.
36 [?]
Bateman, MHD Instabilities (1980) [?]
Biskamp, Nonlinear Magnetohydodynamics (1993) [?]
The neoclassical MHD model for axisymmetric toroidal plasmas is described in
J.D. Callen, W.X. Qu, K.D. Siebert, B.A. Carreras, K.C. Shaing and D.A.
Spong,Neoclassical MHD Equations, Instabilities and Transport in Tokamaks,
in Plasma Physics and Controlled Nuclear Fusion Research 1986 (IAEA, Vienna,
1987), Vol. II, p. 157 [?]

PROBLEMS
6.1 Use the definition of the pressure in (??) with vr v V to show that the
isotropic pressure of a species in the center-of-mass frame (V) of an MHD plasma
is
pCM
= ps + (ns ms /3)|Vs V|2. /
s
6.2 Show that the plasma momentum equation (6.6) obtained by adding the electron
and ion momentum equations is exact (i.e., it does not involve an me /mi << 1
approximation). [Hint: To obtain the inertia term on the left it is easiest to use
(??) for the electron and ion momentum equations. Also, first show that

!
s

ms ns Vs Vs = m VV +

!
s

ms ns (Vs V)(Vs V)

in which V is the MHD mass flow velocity defined in (6.19).] //*


6.3 Evaluate JD and show that it is equal to J , and to the terms on the
right of (6.16). Explain the physical significance of the equality of these two
quantities. //*

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

32

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

6.4 Multiply the electron and ion momentum balance equations (??) by qe /me and
qi /mi and add them to obtain the exact generalized Ohms law
J
+ (JV + VJ q VV) = #0 p2
t

E + VB

&

J"
J
+
"

CM
(1 Zi me /mi )JB [PCM
e Zi (me /mi )Pi ]

(1 + Zi me /mi )ne e

'

in which PCM
ps I + s + ns ms (Vs V)(Vs V) is the pressure tensor of
s
a species in the center-of-mass frame (V) of the plasma. Show that this result
simplifies to (6.17) for me /mi << 1 and strongly subsonic relative species flows
(|Vs V|/vT s << 1). [Hint: Use ne /ni = qi /e = Zi for this two species plasma
and
mi ni (J/ne e)
me ne (J/ne e)
Ve = V
, Vi = V +
.] ///
m e ne + m i ni
me ne + mi ni
6.5 Show that the electron inertia term is negligible compared to the electric field
in the parallel generalized Ohms law for kc/pe << 1. [Hint: Use the parallel
component of the nonrelativistic Amperes law: 2A" = 0 J" from (??).] //*
6.6 Show that for a wavelike perturbation in a sheared slab model magnetic field
the perturbed electron pressure gradient is negligible in the parallel generalized
Ohms law when (6.19) is satisfied and >> e . [Hint: When the magnetic
& B0 + A
" b
and " b

field is perturbed in MHD B B0 + A
is changed accordingly.] //*

6.7 Show that the perpendicular electron inertia term is a factor of at least /ce
smaller than E in (6.20) and hence negligible in MHD. [Hint: Show that for the
diamagnetic and polarization MHD currents the electron inertia term is smaller
than that due to the electron polarization flow (??).] //*
6.8 Derive the MHD system momentum density equation (6.35). [Hint: Rewrite the
momentum equation (6.27) using Amperes law and vector identities (??), (??)
and (??).] //
6.9 Derive the MHD system energy density equation (6.36). [Hint: Take the dot
product of V with the MHD momemtum equation (6.45), and simplify the result
using Ohms law in the form VB = J E, vector identities (??) and (??),
and
1 P

VP =
+
P V J 2,
1 t
1
which is obtained from a combination of the equation of state (6.29) and the
mass density equation (6.25).] //
6.10 Use the tensor form of Gauss theorem (??) to calculate the force on a volume
of MHD fluid in terms of a surface integral over the stress tensor. Use an
infinitesimal volume form of your result to discuss the components of the force
directions. //
in the
ex ,
ey , b
6.11 The pressure tensor in an open-ended magnetic mirror is anisotropic because
of the loss-cone. a) Show that for species distribution functions fs which do
not depend on the gyroangle the pressure tensor is in general of the form
b)
+ P" b
b
in which P = P (, , B) and P" = P" (, , B). b)
P = P (I b

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

33

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

Work out P. c) Show that the condition for force balance along a magnetic
P = 0) can be reduced to
field (b

P" ::
P" P
=
.
B :,
B

d) Discuss how this result indicates confinement of plasma along the magnetic
field in a magnetic mirror. ///
6.12 Obtain the angle between J and B in a screw pinch equilibrium as a function
of a relevant plasma . //*
6.13 Consider a pressure profile given by P (x)/P (0) = exp(x2/a2 ) in a sheared slab
magnetic field model with no curvature or shear. a) Calculate the diamagnetic
current. b) Determine the Bz (x) profile induced by this diamagnetic current.
c) Show that the plasma pressure produces a diamagnetic effect. d) Show that
your Bz (x) agrees with (6.59). //
6.14 Consider the MHD radial force balance equilibrium of a cylindrical plasma with
a pressure profile P (r) that vanishes for r a which is placed in a uniform
magnetic field B = B0
ez . Use the azimuthal component of Amperes law for J
and solve the resultant force balance equation for Bz (r). Show that your result
agrees with (6.59). /
6.15 One definition of a magnetic field line is d#B = 0. Show that its time derivative yields the magnetic evolution equation (6.71). How does this show that a
magnetic field line is advected with the moving plasma in the ideal MHD limit?
[Hint: Use vector identities (??) for (d/dt)d# and (??), (??).] //
6.16 Show that for a Clebsch magnetic field representation B = the ideal
MHD evolution equation (6.71) is satisfied by d/dt = d/dt = 0. Why does
this show that a magnetic field line is advected with an ideal MHD plasma? //
6.17 Derive the canonical flux invariant for an isentropic plasma species that is a
combination of the magnetic flux and species vorticity flux which is deduced
from the canonical momentum (??) ps = ms v + qs A as follows. a) First,
s =
average the canonical momentum over a Maxwellian distribution to obtain p
ms Vs + qs A. b) Next, use this result to define a species canonical flux invariant
#s

))

&

dS A +

ms
Vs
qs

'

))

&

dS B +

'

ms
Vs .
qs

c) Obtain d#s /dt and use the species momentum equation (??) to show that
d#s
=
dt

))

dS

&

ps + s Rs
ns q s

'

d) Show that d#s /dt = 0 for an isentropic plasma species. e) Discuss how the
canonical flux invariant #s combines the ideal MHD frozen flux theorem (6.73)
and the Kelvin circulation theorem (??). f) Indicate the physical processes
that can cause net transport of a plasma species relative to the canonical flux
surfaces #s . g) Why doesnt inertia contribute to transport relative to the #s
surfaces? [Hint: Use vector identities (??), (??) and (??) in part c).] ///*

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

34

CHAPTER 6. PLASMA DESCRIPTIONS II: MHD

6.18 Show that for the MHD model the electron and ion canonical fluxes defined in
the preceding problem are, to lowest order in (me /mi )1/2 << 1,
#e &

))

&

dS 1

c2 2

2
pe

'

B,

#i &

))

&

dS B + B

c
V

pi
cA

'

Use these two relations to discuss the degree to which the magnetic field is frozen
into the electron and ion fluids in an ideal MHD plasma. //*
6.19 Show that the total mass M of an MHD plasma in a volume V that moves with
the plasma flow velocity V will be conserved if the mass density satisfies the
mass density equation (6.25). [Hint: Determine the condition for dM/dt = 0
and use vector identity (??) for (d/dt)d3x.] //
is spatially
6.20 Work out the terms on the right of (6.82). [Hint: Since B = B0 b

uniform, it commutes with the " b and = b" operators.] //


6.21 Work out formulas for the ratio of the electron and ion thermal speeds to the
Alfven speed in terms of e 20 pe /B 2 and i 20 pi /B 2 . What are these
ratios for a = 0.08, Te = Ti , electron-deuteron plasma? /

6.22 How large would the magnetic field strength B have to be for the Alfven speed
to be equal to the speed of light for ne = 1020 m3 and Ai = 2? /
6.23 Show that for perturbations on the equilibrium time scale for compressional
Alfven waves
" /B0 = (/2) (P /P0 ). /
B
6.24 Since to lowest order in me /mi << 1 the MHD momentum equation results from
the ion momentum equation, on the equilibrium time scale for compressional
Alfven waves the radial component of the ion momentum equation should be in
equilibrium. Show that the equilibrium radial ion momentum (force) balance
equation in a screw pinch plasma yields the following relation for the axial flow
in terms of the radial electric field, pressure gradient and poloidal flow:
Viz

1
=
B

&

'

d0
1 dp0i
+
Vi Bz . /
dr
n0i qi dr

6.25 Determine the frequency ranges where an electromagnetic wave impinging on


an unmagnetized plasma: a) propagates through it, b) is evanescent on a c/pe
length scale, and c) dissipatively decays in a resistive skin depth (6.94)? [Hint:
review Section 1.5 and consider a time-dependent electrical conductivity.] //
6.26 Show that the Lundquist number can be written in terms of fundamental microscopic variables as
ce a
a
S=
e c/pi L"
Should S always be a large number for a magnetized MHD plasma? /

DRAFT 10:31
January 28, 2003

c
!J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

Appendix A

Physics Equations, Units,


and Constants
This appendix provides a summary of the fundamental physical laws from other
areas of physics, as they are commonly used in plasma physics. Key equations,
units and physical constants are given for mechanics, electrodynamics, statistical
mechanics, kinetic theory of gases, stochastic diffusion processes, fluid mechanics and quantum mechanical effects. While the procedures for deriving these
equations are given in outline form, details are omitted. Readers should consult the textbook references listed at the end of each section for more detailed
explanations and theoretical developments. In some parts of this appendix
extensive use is made of the vector algebra and calculus relations given in Appendix D. The International System of Units (Syst`eme International dUnites),
often called mks units, are used throughout this appendix, and the book. Physical constants and SI unit interrelationships are given in tables in Section A.8
at the end of this Appendix.

A.1

Mechanics

Newtons second law states that the mass m times the acceleration a of a particle
is given by the force F (in units of newtons or kg m/s2 )
ma = F,

Newtons second law.

(A.1)

A conservative force is one that is derivable from the gradient of a potential


that is independent of time:
F = V (x),

conservative force.

(A.2)

Since the acceleration in (A.1) is just the time derivative of the particle velocity,
a dv/dt, taking the dot product of the velocity v with Newtons second law
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS


for a conservative force yields:
!
"
d mv 2
+ V (x) = 0
=
dt
2

mv 2
+ V (x) = constant,
2
energy conservation,

(A.3)

where mv 2 /2 m(v v)/2 T is the particle kinetic energy and V (x) is the
potential energy. The SI unit of energy is the joule (J), which is equal to a
newton meter (N m). In plasma physics particle energies are usually quoted
in electron volts (eV), which is the energy in joules divided by the elementary
charge [ eV J/e = J/(1.602 1019 ) ].
The force on a particle of charge q subjected to an electric field E(x, t) and
a magnetic induction field B(x, t) is
F = q (E + vB) ,

Lorentz force.

(A.4)

For electrostatic situations with no magnetic field, the electric field can be written in terms of the electrostatic potential (x) : E = . Then, the Lorentz
force becomes conservative [see (A.2)] with V (x) = q(x), and the energy
conservation relation (A.3) is applicable.
When only a magnetic field is present, the combination of Newtons second
law and the Lorentz force becomes
dv
dv
= qvB
=
= c v,
(A.5)
m
dt
dt
where
c qB/m,

the angular velocity,

(A.6)

for gyromotion of the charged particle in the magnetic field. The negative sign
is needed in this vectorial definition so that charged particles gyrate according
to the right-hand rule with the thumb pointing in the direction of c . The
magnitude of c gives the radian frequency (rad/s) for the gyromotion:
c = qB/m,

gyrofrequency,

(A.7)

which is also called the cyclotron (the source of the subscript c) or Larmor1
frequency. This formula is unchanged for relativistic particles except
# for the
fact that then the mass becomes the relativistic mass: m m/ 1 v 2 /c2 .
Since the dot product of (A.5) with the velocity v vanishes, the particle kinetic
energy is constant a magnetic field does no work on a charged particle in its
gyromotion. In gyromotion a charged particle executes a circular motion about
the magnetic field B with a radius of
# v /c ,

gyroradius,

(A.8)

in which v is the magnitude of the velocity component perpendicular to the


magnetic field direction [v B(Bv)/B 2 ].
1 Actually,

the Larmor frequency is defined to be half the cyclotron frequency.

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

For situations where both electric and magnetic fields are present, it is convenient to write them in terms of the scalar potential (x, t) and vector potential
A(x, t): E = A/t, B = A see (A.55). Then, Newtons second
law (A.1) for a nonrelativistic charged particle subjected to the Lorentz force
(A.4) can be written as
$
%
U
d U
dv
=
+
, U = q q(v A)
(A.9)
m
dt
x
dt v
in which U is a generalized potential energy, U/x U , U/v v U ,
where v is the gradient in velocity space. The single particle Lagrangian, which
has units of energy and is given by the kinetic energy minus the generalized
potential energy, is defined by
L(x, v, t) T U =

mv 2
q + q(v A),
2

Lagrangian,

(A.10)

where again T = mv 2 /2 is the particle kinetic energy. The vector equation of


motion for a charged particle (i.e., Newtons second law with the Lorentz force)
can be written in terms of the Lagrangian as
&
' (
'
L ''
d L ''
= 0, Lagrangian equations of motion.
(A.11)

dt v '
x '
t,x

t,v

For an orthogonal coordinate system with unit base vectors


ek , the orthogonal
projections of this vector equation yield
$
%
d L
L
= 0, k = 1, 2, 3 Lagranges equations.
(A.12)

dt qk
qk

ek x and the velocity coordinates are


Here, the spatial coordinates are qk
ek v =
ek dx/dt. Note that, like Newtons second law, Lagranges
qk
equations are in general second order ordinary differential equations in time.
It is often convenient to change the charged particle equation of motion into
two coupled first order differential equations. To effect this change one first
defines
L
= mv + qA, canonical momentum,
(A.13)
p
v
in which v dq/dt. Next, the single particle Hamiltonian function H, which
also has units of energy, is defined through the Legendre transformation:
H(x, p, t) p

|p qA|2
dx
L =
+ q T + V,
dt
2m

Hamiltonian. (A.14)

It is the sum of the kinetic energy (T = |p qA|2 /2m = mv 2 /2) and the
potential energy (V = q), and by construction is independent of velocity:
H/v|x,p,t = 0. The equation of motion for a charged particle can be written
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

(for both orthogonal and nonorthogonal coordinate systems) in terms of the


Hamiltonian function H, as two coupled first order vector differential equations
in time:
'
'
H ''
dx
H ''
dp
=
,
=
, Hamiltons equations of motion. (A.15)
dt
x 't,p
dt
p 't,x

The total time derivative of the Hamiltonian is given via chain-rule partial
differentiation by
'
'
'
H ''
dx H ''
dp H ''
dH
=

+
+
.
dt
t 'x,p
dt x 't,p
dt p 't,x

Using Hamiltons equations of motion, the sum of all terms except the explicit
partial time derivative vanish because the Hamiltonian does not vary along
the charged particles motion in the relevant (x, p) six-dimensional phase space:
dx/dt H/x + dp/dt H/p = 0. Thus, the total time derivative of the
Hamiltonian is simply
'
'
$
%
H ''
A
L ''

dH
=

,
(A.16)
=

=
q
dt
t 'x,p
t 'x,v
t
t

which indicates the increase in energy due to a temporally increasing potential


and due to the work v qE done by the inductive component A/t of the
electric field.
Projecting out the geometrical components of the Hamiltonian form of the
equations of motion (A.15) for a charged particle in the orthogonal directions

ek yields
H
dpk
=
,
dt
qk

dqk
H
=
,
dt
pk

k = 1, 2, 3,

Hamiltons equations,

(A.17)

in which pk
ek p are the canonical momentum coordinates and qk
ek x
are the conjugate spatial coordinates.
The various equations of motion have been written in forms that are independent of the coordinate system and they are valid in the initial coordinate
system as well as transformed ones. Also, the equations are valid for nonrelativistic particles (v << c) and are all Galilean invariant. That is, they are
unchanged upon transformation to another inertial (non-accelerating) frame according to v# = v + Vf and E# = E + Vf B, where Vf is the velocity of the
second inertial frame (subscript f ) relative to the first.
When the potential and vector potential A do not depend explicitly on
time, (A.16) shows that the Hamiltonian is a constant of the motion:
H =

mv 2
|p qA|2
+ q =
+ q = = constant,
2m
2
A

= 0,
= 0, energy conservation.
for
t
t

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

(A.18)

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

In such conservative systems the particle energy and time are canonical conjugate Hamiltonian coordinates: p = and q = t.
When the gradient of the Hamiltonian vanishes in a particular direction ek ,
(A.15) shows that the conjugate canonical momentum in the same direction is
a constant of the motion:
H
= 0,
x
canonical momentum conservation.

pk ek p = ek (mv + qA) = constant for ek

(A.19)

In an orthogonal coordinate system the base vector ek becomes the unit vector

ek ; then, the criterion for canonical momentum conservation in the


ek direction
becomes simply H/qk = 0 (i.e., H independent of the coordinate qk , which
implies symmetry in the
ek direction).
Lagranges or Hamiltons equations of motion can be derived (by considering
variations with x and v dx/dt or p as the independent variables, respectively)
from Hamiltons variational principle of least
action (time integral of difference
)
between kinetic and potential energy): L dt = 0. It can also be shown using
(A.14) that for a conservative system
where the Hamiltonian is a constant of the
)
motion [see (A.18)], the action p dq is a variational quantity along a particle
trajectory.
For periodic motion in a given coordinate qi it is convenient to introduce as
a variable the action integral over a cycle:
*
+ 2
1
1
qi
pi dqi =
di pi
, action variable.
(A.20)
Ji =
2
2 0
i
The action variable, which is a momentum-like quantity, is the area in pi , qi
phase space encompassed by the periodic motion. The canonically conjugate
action-angle i is the angular or cyclic variable corresponding to periodic motion
around the perimeter of this area. Hamilton-Jacobi theory (see references at end
of this section) can usually be used to determine the action angle coordinate
qi i . Writing the Hamiltonian in terms of the action variable Ji , the Hamilton
equation dqi /dt = H/pi [see (A.17)] becomes:
H
di
=
i ,
dt
Ji

action angle evolution equation.

(A.21)

The period of the oscillatory motion can be determined in general from


*
*
*
dqi
dqi
=
, oscillation period.
(A.22)
i dt =
dqi /dt
H/pi

The radian frequency for the periodic motion is


i (Ji ) = 2/i ,

oscillation frequency.

(A.23)

The Hamiltonian for a periodic system in action-angle variables is thus simply


Hi = i Ji ,
DRAFT 11:16
September 2, 2003

action-angle Hamiltonian.

c
#J.D
Callen,

(A.24)

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

For nearly periodic motion in situations where the generalized potential U


in (A.9) varies slowly and aperiodically in space and time (compared to the
oscillations), the action in (A.20) is nearly constant and given by the ratio of
the oscillation energy to the oscillation frequency:
Ji '

Hi
i
=
,
i
i

action for nearly periodic motion.

(A.25)

For slow, temporal changes that are characterized by a parameter a(t) and are
not themselves periodic, it can be shown that, while the slow variations cause
linearly small, oscillatory [ (a/
i a) sin i t << 1] changes in J, the average
of dJ/dt over an oscillation period is quadratically small in the rate of
temporal change:
.$
/
%2
,
*
di dJi
a
Ji
1
a

dJi
= 0+O
,
(A.26)

, 2
dt i

i a
i a
i dt
where the dots over quantities indicate their time derivatives. For such situations the action Ji is called an adiabatic invariant; it is often a very useful
approximate constant of the motion. When the#small variations in the potential oscillate at a slow frequency i >> a (1/a) 2a/t2 , harmonics of
this slower oscillation that are resonant with the fundamental oscillations (i.e.,
i = n a , n an integer) can lead to secular changes in the action Ji that grow
slowly in time. Hence they can break the constancy of the adiabatic invariant
over a long time period. This usually occurs when the slow oscillations exceed a
small critical amplitude (typically 0.1 of the main oscillations). The relevant
multiple time scale analysis and conditions for such breakdowns of adiabaticity
are discussed in E.6.
As an example of the use of mechanics theory, consider the central-force
problem of determining the scattering angle and elastic cross-section for a
Coulomb collision of two non-relativistic, charged particles. Assume a charged
particle of species s with charge qs , mass ms and initial velocity v experiences a
Coulomb collision (i.e., interaction via the Coulomb electric field force) with another charged particle of species s# with parameters qs! , ms! and v# . Multiplying
the force balance equations obtained from (A.1) and (A.4) for each particle by
the mass of the other particle and subtracting, taking account of the equal and
oppositely directed electric field forces on the two particles due to the Coulomb
potential [qE = q, = q/({4+0 }|x|) see (A.33)], yields the equation
of motion for the two-particle system in the center-of-momentum coordinate
system:
mss!

q s q s!
1
d
(v v# ) =

dt
{4+0 } |x x# |

(A.27)

in which mss! = ms ms! /(ms + ms! ) is the reduced mass for the two particle
system.
Initially, when the particles are very far apart, one can define the impact
speed as u = |v v# | and the collision impact parameter as b (distance of closest
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

approach if the Coulomb electric field did not deflect the particles). Further,
one defines the classical (i.e., not quantum mechanical) minimum distance of
closest approach as
bcl
min

q s q s!
,
{4+0 }mss! u2

classical minimum impact parameter

(A.28)

at which the center-of-momentum kinetic energy is half the Coulomb potential

energy [qs qs! /({4+0 }bcl


min )] and below which large-angle deflections (> 90 ) can
be expected to occur.
Since the collision takes place in a plane defined by the vectors x x# and
v v# , it is convenient to define instantaneous radial and angular coordinates
in the center-of-momentum frame by the radial separation of the particles r
|x x# | and by the angle that the line x x# makes with the line |x x# |
when the particles were initially very far apart. In these coordinates the angular
momentum p [constant because of symmetry of the Coulomb potential in the
direction see (A.19)] and total energy [constant because the potential
does not depend explicitly on time see (A.18)] can be written as
p = mss! r2 = mss! bu,
0
1
1
1
q q!
= T + V = mss! r 2 + r2 2 + s s = mss! u2 .
2
{4+0 }r
2

Solving the second (energy conservation) equation for r and dividing by the
obtained from the first equation yields
2
r
dr
2
=
r2 2 r bcl
min b ,
d
b

where the sign is negative when the particles are approaching each other and
positive as they recede.
At the minimum or closest approach distance rm , dr/d = 0. The angle m
at this point is given by
$ cl %
+
b

dr
= arctan min + .
m =
b
2
rm dr/d

In the center-of-momentum frame the angular deflection caused by the collision is given by 2m and hence
tan

bcl
qs q s !

= min =
,
2
b
{4+0 }mss! u2 b

scattering angle.

(A.29)

Thus, b > bcl


min causes Coulomb scattering by less than 90 ( < /2), while
cl

b < bmin induces more than 90 scattering.


The differential cross-section d (measured in meters2 or barns 1028 m2 )
by which Coulomb collisions of incoming charged particles of species s with
impact parameter b and azimuthal angle scatter off of charged particles of

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

species s# into spherical angles , within the differential solid angle d


sin d d is thus given [using (A.29) to write b() = bcl
min / tan(/2)] by
' '
$
%2
bcl
b '' db ''
min
d
=
d
d = b db d =
sin ' d '
2 sin2 /2

or,

d
=
d

%2
%2
$
bcl
q s q s!
min
=
,
2 sin2 /2
2{4+0 }mss! u2 sin2 /2
Rutherford differential scattering cross-section.

(A.30)

Standard intermediate level mechanics textbooks, which include extensions


to relativistic systems, are:
Symon, Mechanics (1971) [?].
Barger and Olsson, Classical Mechanics: A Modern Perspective (1973) [?].

The standard advanced level mechanics textbook is:


Goldstein, Classical Mechanics (1950, 1980) [?].

A.2

Electrodynamics

An electrostatic theory is appropriate for time-independent charge density distributions q (x), electric fields E(x), and magnetic induction fields B(x). In electrostatics the irrotational (E = 0) electric field E with units of volts/meter
is written in terms of the scalar potential (x) with units of volts, E ,
and related to the charge density distribution:
Gauss law,

(A.31)

Poissons equation.

(A.32)

E = q /+0 ,
2 = q /+0 ,

The charge density distribution has units of coulombs/meter3 . For localized


charge density distributions [lim|x| q (x) 0] the general (Green-functiontype) solution of Poissons equation in an infinite medium is [see also (??)]
+
q (x# )
.
(A.33)
(x) = d3x#
{4+0 } |x# x|
For a point charge at x = x0 the charge distribution is q (x) = q (x x0 ) and
the potential becomes
(x) =

q
,
{4+0 } |x x0 |

Coulomb potential.

(A.34)

Here and throughout this book the mks factor {4+0 } is written in braces;
eliminating this factor yields the corresponding cgs forms of these electrostatic
response formulas.
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

In a dielectric (ponderable) medium the charge density q is composed of a


part free due to free charges and a part due to a polarization charge density,
pol = P where P +0 E E is the presumed linear and isotropic polarization (units of coulomb/meter2 ) of the medium induced by the electric field E
and E is the dimensionless electric susceptibility of the medium:
q = free + pol = free P = free +0 E E.

(A.35)

Thus, in an isotropic dielectric medium Gauss law becomes Coulombs law:


D = free ,

D +0 E + P = +0 (1 + E ) E = + E

(A.36)

in which the mediums dielectric constant + +0 (1 + E ) is the constituitive


relation between the displacement vector D and the electric field E.
A magnetostatic theory is appropriate for time-independent current density
distributions J(x) and magnetic induction fields B(x). In magnetostatics the
solenoidal (transverse, B = 0) magnetic induction field B which has units
of weber/meter2 or tesla can be written in terms of the vector potential A, i.e.,
B = A, and related to the current density distribution J which has units
of ampere/meter2 :
B = 0 J

2 A = 0 J,

static Amperes law

(A.37)

in which A = 0 (the Coulomb gauge) has been assumed in the last form. For
localized current density distributions [lim|x| J(x) = 0], the general (Greenfunction-type) solution for A in an infinite medium is [see also (??)]
3 4+
J(x# )
0
d3x# #
.
(A.38)
A(x) =
4
|x x|

The magnetic field around an infinite wire carrying a current I (amperes) along
the z axis of a cylindrical coordinate system can obtained from this equation
using B = A and a current density J =
ez (I/2r) lima0 (r a):
B =

0 I

e ,
2r

magnetic field around a current-carrying wire.

(A.39)

To obtain the corresponding cgs form of this and other magnetic field response
equations, eliminate the {0 /4} factor and replace J by J/c, or replace 0 J
by 4J/c.
In a magnetizable medium the current density J is composed of a part Jfree
due to the current induced by free charges and a magnetization current density Jmag = M where M M H is the presumed linear and isotropic
magnetization (units of ampere-turns/meter) of the medium induced by the
magnetic field H (units of ampere-turns/meter) and M is the dimensionless
magnetic susceptibility of the medium:
J = Jfree + Jmag = Jfree + M = Jfree + (M H).
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

(A.40)

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

10

Thus, in an isotropic, magnetizable medium the static Amperes law becomes


H = 0 J,

H = B/0 M,

B = 0 (1 + M ) H = H

(A.41)

in which = 0 (1 + M ) is the magnetic permeability of the medium and


B = H is the constituitive relation between the magnetic induction B and the
magnetic field H.
The microscopic Maxwell or electromagnetic (em) equations for determining
time-varying electric and magnetic fields caused by charge and current density
distributions q (x, t) and J(x, t), respectively, in free space are
Maxwell Equations
name
differential form

integral form

Gauss law
E =

q
+0

Faradays law
E =

B
t

+
++
q
# dS E =
d3x
+0
S
V

(A.41a)

(A.41b)

no magnetic monopoles
B = 0

d" E =

++

dS B

++
# dS B = 0

(A.41c)

Amperes law $
%
E
B = 0 J + +0
t

d" B = 0

++

E
dS J + +0
t
S

(A.41d)
(A.42)

Here, +0 is the electric permittivity of free space which has units of farad/meter
= coulomb/(volt meter) = joule/(volt2 meter), 0 is the magnetic permeability of free space which has units of henry/meter = weber/(ampere meter) =
weber2 /(joule meter) and 0 +0 = 1/c2 where c is the speed of light in free space.
Taking the divergence of Amperes law and making use of Gauss law yields
q
+ J = 0,
t

continuity equation for charge and current.

(A.43)

In Amperes law the displacement current +0 E/t was introduced by Maxwell


to make the electrodynamics equations consistent with the charge and current
continuity equation (A.43).
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

11

The physical significance of the source


) 3 terms on the right of the integral
d x q Q, net charge (in coulombs)
forms of the Maxwell equations
are:
V
))
dS

,
magnetic
flux (in webers) penetrating
within the volume
V
;
S
))
the surface S; S dS J I, the total electric current (in amperes) flowing
through the surface S. The Maxwell equations are relativistically invariant; in
particular, they are invariant under Lorentz transformations, which preserve the
constancy of the speed of light, independent of the motion of the source, upon
transformation to another inertial rest frame.
The corresponding macroscopic Maxwell equations in an isotropic, polarizable, magnetizable medium are written in terms of D + E = +0 E + P and the
magnetic field H B/0 M, and free charge, current densities free , Jfree :
D = free ,

E = B/t,

B = 0,

H = Jfree + D/t.
(A.44)

The total rate at which


) the electromagnetic (em) fields do work on a medium
in a finite volume V is V d3x Jfree E the magnetic field does no work since
the magnetic force qvB on charged particles is perpendicular to the velocity.
Using the macroscopic Maxwell equations to calculate the rate of doing work
yields the energy conservation law for electromagnetic fields:
wem
+ Sem = Jfree E,
t

em field energy conservation,

(A.45)

where
wem wE + wB
Sem = EH,

1
(E D + B H),
2

em energy density (J/m3 ),

Poynting vector (flux of em energy) (J/m2 s),

Jfree E = joule heating (W/m3 = J/m3 s = VA/m3 ),

(A.46)

(A.47)

(A.48)

in which the energy densities in the electric and magnetic fields are defined by
wE

1
(E D),
2

electric field energy density (J/m3 ),

(A.49)

wB

1
(B H),
2

magnetic field energy density (J/m3 ).

(A.50)

A corresponding momentum conservation equation for electromagnetic fields


can be deduced, for situations where charge and current densities are present in
free space, from the microscopic Maxwell equations:
gem
+ Tem = q E + JB,
t
DRAFT 11:16
September 2, 2003

em field momentum conservation,

c
#J.D
Callen,

(A.51)

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

12

where
gem
Tem = +0

1
EB,
c2

momentum density in em fields,

"
"
!
1 |B|2
|E|2
I EE +
I BB
2
0
2

em stress tensor,

q E + JB = momentum input to em fields from medium.

(A.52)

(A.53)

(A.54)

Here, the electromagnetic stress tensor is defined to be opposite in sign from


the usual Maxwell stress tensor in electrodynamic theory [see Eq. (6.119) in
Jackson, Classical Electrodynamics, 3rd Edition (1999)[?]] so that the electromagnetic stress can be added to the pressure tensor P to obtain the total force
density in a plasma in the form F = (P + Tem ). For a dielectric medium
the conservation of momentum for electromagnetic fields depends somewhat on
the medium considered because of the possible ambiguity as to which parts of
q E + JB belong to the dielectric and which parts represent free charge and
current densities.
Since the magnetic induction field B is a solenoidal or transverse field ( B
= 0), it can be represented in terms of a vector potential A, i.e., B = A.
Using this representation, Faradays law can be written as (E+A/t) = 0,
which indicates that E + A/t can be represented in terms of the gradient of a
scalar potential . Thus, the electromagnetic fields E and B can be represented
in terms of the potentials (units of volts) and A (units of weber meter):
E = A/t,

B = A,

em fields in terms of potentials. (A.55)

In terms of the potentials , A the inhomogeneous, microscopic Maxwell equations (Gauss and Amperes laws) become (0 +0 = 1/c2 )
$
%
%
$
1 2
1 2
q
2
2
2 2 A = 0 J,
(A.56)
2 2 = ,
c t
+0
c t
in which
A +

1
= 0,
c t

Lorentz gauge condition,

(A.57)

which provides a constraint relation between the potentials, has been used. [If
the Coulomb gauge (A = 0) is used, the equations (A.56) are different.]
For a dielectric medium [i.e., a medium that is polarizable (q = P)
and magnetizable (J = M) but not significantly conducting which would
imply J = E], equations (A.56) become scalar wave equations of the form
%
$
2
2 + 2 u(x, t) = S(x, t), dielectric medium wave equation. (A.58)
t
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

13

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS


Sinusoidal plane wave solutions of this equation are in general of the form
ei ,
u(x, t) = u
k, ei(kxt) u

Fourier plane wave Ansatz,

(A.59)

where by convention u
is a complex constant for a given k, and physical waves
u(x, t) are obtained by taking the real part: u(x, t) Re{
u ei }. Substituting
this Ansatz into the sourceless (S = 0) wave equation yields the dispersion
relation (relationship between and k) for nontrivial (normal mode) solutions:
2 =

k2
k 2 c2
= 2
+
n

in which
n

=
ck
,

kc
,
n

light waves in a dielectric,

index of refraction.

(A.60)

(A.61)

The index of refraction is the ratio of the speed of light in vacuum to that in
the medium.
For a given k, , a point of constant wave phase in u(x, t), which is defined
by 0 = d/dt = k dx/dt k V , moves at
V

ek ,
k

wave phase velocity,

(A.62)

in which
ek k/k is the unit vector along k. The phase velocity for light waves
in a dielectric medium is the speed of light in the medium in the direction of
wave propagation (k): V = (c/n)
ek . Since a steady, monochromatic (single
k, ) carrier wave carries no information, the wave phase speed can be greater
than the speed of light. A wave packet, which results from superposing waves
of different k, , carries information at
Vg

= k (k),
k

wave group velocity,

(A.63)

whose magnitude must, by causality, be less than or equal to the speed of light.
For nondispersive media [n/k = 0 = n = n()], the group velocity is the
same as the phase velocity. Thus, the group velocity of light waves in typical
(nondispersive) dielectric media (e.g., water for visible light) is the same as their
phase velocity. Since plasmas are typically dispersive media for ranges of k, of
interest, the group velocities of waves in plasmas are often different from their
phase velocities.
The electric field for the most general homogeneous transverse (kE = 0)
plane wave propagating in the direction k can be represented by
E(x, t) = (#1 E1 + #2 E2 ) ei(kxt) ,

polarization representation.

(A.64)

Here, #1 , #2 are mutually orthogonal wave polarization unit vectors in direcek ) and
tions perpendicular to the direction of wave propagation (#1 #2
E1 , E2 are in general complex numbers. If E1 and E2 have the same complex
phase, the wave is linearly polarized. If E1 and E2 have the same magnitude, but
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

14

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

differ in phase by 90 degrees the wave is circularly polarized. A representation


that is useful for circularly and elliptically polarized waves is
E(x, t) = (E+ #+ + E # ) ei(kxt) ,

alternative representation,

(A.65)

in which E+ and E are complex amplitudes and


1
# (#1 i#2 ),
2

rotating polarization unit vectors.

(A.66)

The E+ |E+ |ei+ term represents a positive angular momentum and helicity
(left circularly polarized in optics2 ) wave that rotates (for decreasing phase
k x t + + at a fixed point in space) in theclockwise direction relative to the k direction since Re{E(x, t)+ } = (|E+ |/ 2)(#1 cos #2 sin ).
Conversely, the E term represents a negative angular momentum and helicity
(right circularly polarized) wave that rotates in the opposite direction. Circularly polarized waves are represented by either E+ or E , depending on whether
they have positive or negative helicity. A wave is elliptically polarized if it has
both E+ and E components and they are dissimilar when E+ /E = 1,
one reverts to a linearly polarized wave.
Standard intermediate level textbooks for electrodynamics, or electricity and
magnetism as it has been called historically, are:
Reitz, Milford and Christy, Foundations of Electromagnetic Theory (1979) [?]
Lorrain, Corson and Lorrain, Electromagnetic Fields and Waves (??) [?]
Barger and Olsson, Classical Electricity and Magnetism: A Contemporary Perspective (1987) [?].

The standard advanced level textbooks are:


Jackson, Classical Electrodynamics (1962, 1975) [?]
Panofsky and Phillips, Classical Electricity and Magnetism (1962) [?].

A.3

Statistical Mechanics

A closed system of particles is in equilibrium in a statistical mechanics sense


if for subsystems thereof all relevant macroscopic parameters are equal to their
mean values to a high degree of accuracy. The particles in a system are weakly
interacting and thus statistically independent if the total system Hamiltonian
is approximately just the sum of the Hamiltonians for the individual particles.
That is, the part of the total system Hamiltonian that represents interactions
between particles must be small, or vanishing, except for infrequent collisions.
2 In optics the rotation direction is determined by the direction of polarization rotation
that would be seen by an observer facing into the oncoming wave. This direction of rotation
is opposite to the modern physics definition which is determined by the direction of rotation
relative to the wavevector k.

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

15

Liouvilles theorem, which follows from the incompressibility [see (A.79)] of


the x, p Hamiltonian phase space for particle trajectories, states that the density of a system of N particles in their 6N -dimensional phase space is constant
along the particle phase space trajectories. A consequence of Liouvilles theorem is that the probability density in the 6N dimensional phase space must
be expressible entirely in terms of constants of the motion. In the macroscopic
rest frame (where the average momentum and angular momentum vanish) of a
system of weakly interacting particles, the only relevant constant (or additive
integral) of the motion is the single particle Hamiltonian.
Statistical mechanics predicts that the most probable distribution of a subsystem of a large number of weakly interacting, free (i.e., monoatomic gas or
unbound) particles in equilibrium with an even larger system of such particles
at a thermodynamic temperature3 T will have a probability density distribution
in the macroscopic rest frame of the system that is given by
(p, q) = 0 eH(p,q)/T ,

Gibbs distribution

(A.67)

in which 0 is a constant and H is the Hamiltonian for a single particle. The


constant 0 is the density of particles
) in)the six-dimensional phase space, which is
obtained from the normalization d3p d3q (p, q) = 1. Thus, for example, the
most probable distribution function for weakly interacting charged particles in
the presence of a potential that is constant in time or slowly varying (compared
to the rate for thermal motion over a relevant scale length for an adiabatic
response, subscript A) is
0 m 13/2
0 m 13/2
2
eH/T = n0
emv /2T q/T ,
(A.68)
fA (x, v) = n0
2T
2T

in which n0 is the equilibrium density (m3 ) of charged particles in the absence


of the potential . The normalization here has been chosen such that integrating
f over the three-dimensional velocity space yields the density distribution
+
(A.69)
nA (x) d3v f (x, v) = n0 eq(x)/T , Boltzmann relation.

This result is applicable for adiabatic processes, i.e., ones that vary slowly compared to the reversible inertial or oscillatory time scales. As an example of
an application of the Boltzmann relation, the gravitational potential near the
earths surface (q V = mgx) confines neutral molecules in the atmosphere
near the earths surface according to the law of atmospheres see (A.137).
In the absence of a potential, (A.68) becomes the Maxwell distribution function:
2
2
0 m 13/2
2
n0 ev /vT
emv /2T =
, Maxwellian distribution,
fM (v) = n0
2T
3/2 vT3
(A.70)
3 Temperatures (and particle energies) in plasma physics are usually quoted in electron
volts, abbreviated eV, and the Boltzmann factor kB that usually multiplies the temperature
T in equations such as (A.67)(A.75) is usually omitted for simplicity.

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

16

Figure A.1: Properies of a Maxwellian distribution function: a) speed dependence; b) number of particles per unit speed v.
#
in which vT 2T /m. The dependence of the Maxwellian distribution on
particle speed v is shown in Fig. A.1a. In spherical velocity-space coordinates
the normalized (by density) integral of the Maxwellian distribution over all
velocity space is [see (??)]
+
+
+
2
fM (v)
4
dv v 2 v2 /vT2
4

=
e
=
dx x2 ex = 1. (A.71)
d3v
n0
0 vT vT2
0
Some of the characteristic speeds
distribution are (see Fig. A.1b):
#
vT = vmax 2T /m,
#

v = 8T /m = (2/ ) vT ,
#
#
vrms = 3T /m = 3/2 vT ,

that can be deduced from the Maxwellian


thermal, most probable speed,
average speed,

(A.72)

root mean square speed.

It is customary in plasma physics to use vT as the reference particle speed since


this is the speed that appears naturally in the exponent of the Maxwellian. This
is the most probable speed because in spherical velocity space the maximum in
2
2
the number of particles with speeds between v and v+dv ( 4v 2 ev /vT ) occurs
at this speed (cf., Fig. A.1b). The average speed v (average of v |v| over the
Maxwellian distribution) is relevant in calculations of the random particle flux
to one side of a plane that is
) whose particles have
) introduced into a medium
a Maxwellian distribution: d3v v 12 | cos | fM = 0 dv v 3 fM = n0 v/4 [see
(??)] where the z axis of the spherical velocity space coordinate system has been
taken to be perpendicular to the plane being introduced. The root mean square
speed vrms (square root of average of v 2 ) is relevant in calculations of the average
kinetic energy mvx2 /2 in a given direction x since all directions are equivalent for

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

17

2
/3):
an isotropic Maxwellian distribution (vx2 = vy2 = vz2 = vT2 /2 = v 2 /3 vrms

$
%
+
2
1
mvx2
mvrms
T
mvx2
3

fM (v) =
= ,
dv
2
n0
2
6
2
one-dimensional particle thermal energy.

The total thermal energy of a particle is given by


$
%
+
2
1
3T
mv 2
mvrms
mv 2

fM (v) =
=
,
d3v
2
n0
2
2
2
three-dimensional particle thermal energy.

(A.73)

(A.74)

Finally, the kinetic pressure embodied in the Maxwellian distribution is


$
%
+
+
2
mv 2
nvrms
3
2
3
fM (v) =
= n T,
p d v (mvx )fM (v) = d v
3
3
kinetic pressure. (A.75)
Entropy is the state of disorder of a closed system. It never decreases
with time: it remains constant for reversible (e.g., Hamiltonian dynamics) processes, but increases for irreversible processes. Irreversible increases in the
entropy of a system are caused by dissipative processes such as the cumulative effects of a large number of random collisions. For a system of weakly
interacting, free particles the entropy is given by the logarithm of the average volume [= 1/0 see (A.67)] of six-dimensional phase space occupied by
a single particle, i.e., s = ln(1/0 ). [For quantum mechanical systems it is
the logarithm of the number of statistically independent states, which is quantized to )be the number of states that fit in the relevant phase space volume:
N qm = d3Np d3Nq eH(p,q)/T /(N ! h3N ) in which h is Plancks constant and
N is the number of degrees of freedom for the system being considered.] Thus,
neglecting constants and using 0 = n0 (m/2T )3/2 n0 /vT3 for an x, v phase
space, for classical systems one has
s = ln (1/0 ) = ln (T 3/2 /n0 ) + constant,

entropy.

(A.76)

For a volume V of uniform density (i.e., n0 = 1/V ) monotonic gas, the entropy
is given by
s = ln V + (3/2) ln T + constant,
which, when multiplied by the molar gas constant R = kB NA is the conventional
form of the entropy for an ideal gas. Alternatively, writing T = p/n0 in (A.76)
so that s = (3/2) ln(pV 5/3 ) + constant, one obtains the constant entropy (isentropic) equation of state pV = constant for an ideal gas in a three-dimensional
system where = 5/3.
Standard intermediate level textbooks for statistical mechanics are:
Kittel, Elementary Statistical Physics (1958) [?]

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

18

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS


Reif, Fundamentals of Statistical and Thermal Physics (1965) [?]
Callen, Thermodynamics (1960) [?]
Kittel and Kroemer, Thermal Physics (1960). [?]

Some advanced level books on statistical mechanics are:


Huang, Statistical Mechanics (1963) [?]
Tolman, The Principles of Statistical Mechanics (1938) [?]
Landau and Lifshitz, Statistical Physics (1959) [?]
Prigogine, Introduction to Thermodynamics of Irreversible Processes (1961). [?]

A.4

Kinetic Theory of Gases

Kinetic theory is a rigorous formalism that is used to provide a description of


the behavior of a large collection of neutral molecules (or atoms) in a gas, particularly when the assumptions of equilibrium statistical mechanics are not valid.
In kinetic theory d3x d3v f (x, v, t) is the (assumed large) number of molecules
located in the six-dimensional (x, v) phase space with spatial positions lying
between x and x + dx and velocity vectors lying between v and v + dv, at
time t. The quantity f (x, v, t), which has units of #/(m3 m3 /s3 ), is called the
distribution function. It is governed by the equation
%
f
F
f
d f (x, v, t)
=
+ vf +
v f =
, kinetic equation (A.77)
dt
t
m
t c
in which F/m is the acceleration of a molecule due to the force F [e.g., the
conservative force in (A.2)], v /v|x,t is the gradient in velocity space,
and f / t)c f represents the effects of abrupt, binary (microscopic)
collisions at rate that result from force fields not included in F.
The (mathematical) characteristics of the first order differential operator (in
the 7 variables x, v, t) on the left of (A.77) represent the trajectories of the
molecules in the absence of collision effects. The first order differential equations governing the trajectories of the particles can be most generally written
using Hamiltons equations. Thus, the kinetic equation for f (q, p, t), where p is
the canonical momentum defined in (A.13) and q is the canonically conjugate
position vector or for f (z, t) where z (q, p) = (x, p) is a six-dimensional variable that represents all of phase space, can be written most generally in terms
of the Hamiltonian variables:
%
f
dq f
dp f
f
dz f
f
d f (q, p, t)
=
+

=
+

=
,
dt
t
dt q
dt p
t
dt z
t c
or, using Hamiltons equations [see (A.15)], as
f
H f
H f
f
d f (q, p, t)
=
+

=
dt
t
p q
q p
t

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

kinetic equation,
(A.78)

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

19

Particle motion in the z (p, q) six-dimensional Hamiltonian phase space is


incompressible:
dx
dp
H
H
dz

= 0,
z dt
x dt
p dt
x p
p x
phase space incompressibility.

(A.79)

Thus, the kinetic equation can also be written in the conservative form
$
%
$
%
$
%
%
f

dq

dp
f

dz
f
d f (q, p, t)
=
+

f +

f =
+

f =
.
dt
t q
dt
p
dt
t z
dt
t c
(A.80)
In the absence of collisions, or for time scales shorter than the collision time,
the solution of (A.77) or (A.78) is that f must be a function of the constants of
the motion see (A.18), (A.19). For collisionless cases where the potentials
, A do not change in time and the Hamiltonian is the only constant of motion,
the solution is f = f [H(p, q)] = f [H(z)]. Assuming further that there are a
large number of molecules which are interacting weakly (e.g., via collisions) with
an even larger number of molecules that have a thermodynamic temperature T ,
and hence that the requirements for the validity of statistical mechanics are
satisfied, the distributions given in (A.67) and (A.68) can be derived from the
kinetic theory of gases.
The microscopic binary collision effects are most generally represented by
%
f
= CB {f (x, v, t)}
t c
+
+
d
3 #
| v v# | [ f (v1 )f (v1# ) f (v)f (v# ) ] ,
dv
d
d
Boltzmann collision operator,
(A.81)
in which v, v# and v1 , v1# are the velocities of the colliding particles before and
after the collision and d/d is the differential scattering cross-section for the
collisions [cf., (A.30)]. Here, for simplicity f (x, v, t) has been written as f (v)
inside the collision operator. In deriving the Boltzmann collision operator it is
assumed that the force F on the left of (A.77) is negligible during the collision
process, that the gas is sufficiently dilute so that binary or two-body collision
processes are predominant (i.e., three-body and many-body collisions or collective particle interactions are negligible), and that the collisions only change the
velocity vectors of the particles (i.e., the collisions abruptly scatter the velocity vectors of the particles at a given point x, t along a particle trajectory).
The Boltzmann collision operator is a bilinear [because of f (v)f (v# )], integral
operator in velocity space. In the absence of radiative effects, since binary collisions conserve particle number, momentum mv and energy mv 2 /2, so does the
Boltzmann collision operator:
+
(A.82)
d3v (v) CB (f ) = 0 for (v) = 1, mv, mv 2 /2.
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

20

The functions (v) are sometimes called summational invariants because linear
combinations of them are also invariants of the collision operator.
For homogeneous (f = 0) gases in equilibrium (f / t = 0) with no
external forces on the molecules (F = 0), the kinetic equation (A.77) becomes
%
f
= CB (f0 ),
(A.83)
0 =
t c
where the subscript zero on f indicates the equilibrium or lowest order solution.
The general solution of this equation is
f0 = fM (v) = n0

m
2T0

%3/2

em|vV0 |

n0 e|vV0 | /vT
,
3/2 vT3
Maxwellian distribution. (A.84)
/2T0

This Maxwellian differs from the statistical mechanics result in (A.70) only by
its
) 3explicit inclusion of the macroscopic flow velocity V0 of the gas (V0
d v vf0 /n0 ), which is not present in (A.70) because that result is obtained
in the rest frame of the gas (i.e., in the V0 rest frame). However, the result is
arrived at by different methodologies in statistical mechanics and kinetic theory.
Kinetic theory provides the more extendable framework for investigating more
complicated situations that do not satisfy the assumptions used in deriving
(A.70) and (A.84).
The Boltzmann collision operator also has the important property of irreversibility: entropy increases until the distribution function is given by (A.84).
taking the entropy functional to be f ln f and defining HB
)Specifically,
d3v f ln f , it can be shown that
+
+
dHB
f
= d3v
(1 + ln f ) = d3v CB (f ) ln f 0,
dt
t
Boltzmann H-theorem,
(A.85)
with the equal sign being applicable only when f becomes equal to the equilibrium, Maxwellian distribution given in (A.84).
In situations close to thermodynamic equilibrium the lowest order distribution is the Maxwellian given by (A.84) and the distortions of the distribution
function are higher order and small. In order to understand he nature of these
distortions and to obtain approximate solutions of the kinetic equation (A.77)
for this situation, consider the expansion of the distribution in a combination of
Laguerre and Legendre polynominials (see Appendix B), which are the complete
2
2
orthogonal basis functions for speed (with the weighting function v 2 ev /vT that
comes from the lowest order Maxwellian distribution in spherical velocity space)

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

21

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS


and spherical angle dependence::
$
!
"
$ %
n (1/2) T (1/2)
v
L0
+
L1
+
P0
moments
f = fM 1 +
n0
T0
vT
5
6
2
(3/2)
(3/2)
+ V1 L1
+
+ 2 v VL0
vT

6
vv (v 2 /3)I 5
(5/2)
(5/2)
:
L
+

L
+

+
1
0
1
2mn0 vT4
+
=

P1

P2

$
$

v
vT
v
vT

%
%

moments
moments

..
.

flmn Ylm (, ) Ln(l+1/2) (v 2 /vT2 ) ev

2
/vT

moment expansion,

(A.86)

lmn

in which the Pl (v/vT ) are Legendre polynomial (spherical velocity space angular) functionals [ 1, v/vT , (vv (v 2 /3)I)/(2vT2 /3) for l = 0, 1, 2 ], the
(l+1/2)
Ln
(x) are (energy functional) Laguerre polynomials with arguments x
mv 2 /2T = v 2 /vT2 , and Ylm (, ) are the usual spherical harmonics that are
proportional to Plm (cos) eim . Useful properties of these special functions are
given in B.5 and B.6. The lowest order parameters of this expansion, which
)
(l+1/2)
moments of the distribution function, correspond
are the d3v Pl (v/v) Ln
physically to: the density (m3 ), flow velocity (m/s) and temperature (eV)
distortions n, V and T away from their equilibrium Maxwellian values of
n0 , V0 and T0 ; the heat flow vector q (W/m2 ), since V1 2q/5nT ; and
the traceless anisotropic part (N/m2 ) of the pressure tensor [see (A.95) below], which has 5 nonvanishing parameters and is sometimes called a kinetic
stress tensor. An approximation in which the moments n, V, T, q and
(= 1 + 3 + 1 + 3 + 5 = 13 moments) are used to represent f is usually called a
Grad 13 moment approximation.
Often one desires a reduced, fluid moment description which integrates the
kinetic equation over velocity space to obtain equations for the physical quantities of density, flow velocity and temperature:
+
n(x, t) = n0 + n d3v f,
(A.87)
density (m3 ):
+
d3v v f /n,
(A.88)
flow velocity (m/s):
V(x, t) = V0 + V
+
temperature (eV):
T (x, t) = T0 + T d3v [m(v V2 )/3] f /n. (A.89)
The relevant fluid moment equations for these quantities are obtained by taking
the relevant velocity-space moments [i.e., the (v) in (A.82)] of the kinetic
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

22

equation in (A.77) using the Boltzmann collision operator in (A.81) and the
conservation properties in (A.82), to obtain
n
+ nV = 0,
(A.90)
t
dV
p ,
= nF
(A.91)
mn
dt
3 dT
n
+ pV = q : V, (A.92)
2 dt

density equation:
momentum equation:
energy equation:
where

+ V,
dt
t

total time derivative,

(A.93)

is the total (partial plus flow-induced advection4 ) time derivative that is some is the average of the single particle
times called the material derivative, and F
force F over a Maxwellian distribution. The higher order moments needed for
closure (complete specification) of these equations are
+
0m
1
|vr |2 f = n T,
p d3v
(A.94)
pressure (N/m2 ):
3
$
%
+
m
5T
|vr |2
f, (A.95)
q d3v vr
conductive heat flux (W/m2 ):
2
2
$
%
+
1
d3v m vr vr |vr |2 I f, (A.96)
stress tensor (N/m2 ):
3
+
= d3v m vr vr f p I P p I
in which vr vV(x, t) is the relative velocity in the frame of reference moving
)
(3/2)
at the macroscopic flow
Note also that q = T d3v vr L1 f. The
) 3velocity V.
2
total heat flux Q d v (m|vr | /2)vr f is the sum of the conductive heat flux
and the convective heat flux: Q q + (5/2)nT V.
The Chapman-Enskog procedure is used to obtain the needed closure relations for collision-dominated situations in which the gas density varies slowly
in space (compared to the collision mean free path v/) and time (compared to the collision time 1/). Then, the lowest order kinetic equation that
describes the distribution function is given by (A.83). Its solution is
f0CE = fM (x, v, t) n(x, t)

m
2 T (x, t)

!
"
|v V(x, t)|2
exp
,
2 T (x, t)
dynamic Maxwellian, (A.97)

"3/2

which is the usual Maxwellian, but now parameterized in terms of the (total)
spatially and temporally varying density, flow velocity and temperature. The
4 In fluid mechanics advection means transport of any quantity by the fluid at its flow
velocity V; convection refers only to the heat flow qconv = (5/2) n T V induced by V.

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

23

conductive heat flux q and anisotropic stress vanish for f0CE . Thus, in order
to determine these needed closure relations, it is necessary to determine the first
order distortion of the distribution function: f f f0CE . (Note that, by
construction, since the total density n flow velocity V and temperature T are
built into f0CE , the density, momentum and energy moments of f vanish:
+
+
+
d3v mv f = 0,
d3v (mv 2 /2) f = 0, C-E consraints.
d3v f = 0,
(A.98)

The kinetic equation for f is obtained by substituting the definition f =


f0CE + f into (A.77), making use of the density, momentum and energy conservation equations to remove the dependences on n/ t, V/ t and T / t.
Neglecting higher order corrections that are inversely proportional to the collision frequency, the result is
!$
$
%
%
$
%"
m|vr |2
5
1
m
|vr |2

vr
T +
W : vr vr
I f0CE ,
CB (f ) '
2T
2
T
T
3
(A.99)
in which, as above, vr v V(x, t), and

9 1
18
V + (V)T I (V) , rate of strain tensor,
(A.100)
2
3
which is caused by gradients in the flow velocity V and has units of per second. The normalized temperature gradient ln T and rate-of-strain tensor W
are called thermodynamic forces because they induce distortions f of the
distribution function away from a dynamic Maxwellian and hence away from
thermodynamic equilibrium. Note that beacause of the invariants of the Boltzmann collision operator given in (A.82), a proper solution of (A.99) for f will
satisfy the Chapman-Enskog constraints in (A.98).
The Boltzmann collision operator needs to be specified in detail in order
to properly solve (A.99). However, the nature of the solution for f can be
exhibited by using an approximate collision model:
W

CK (f ) = f (f f0CE ),

Krook-type collision operator


(A.101)

in which
n v,

collision frequency,

(A.102)

where the overbar indicates the reaction rate v has been averaged over a
Maxwellian distribution. Using this collision operator in (A.99), solving for
f and using the definitions in (A.95) and (A.96) yields the needed closure
(constituitive) relations for the fluid moment equations (A.90)(A.92):
q = m T,

m n m ,

DRAFT 11:16
September 2, 2003

m =

5
5 vT2
= 2 ,
4
4

c
#J.D
Callen,

conductive heat flux,


(A.103)

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS


which is in the form of a Fourier law for the heat flux and
1
v2
1
= 2 m W, m = nm T = nm 2 , viscous stress,
2

2
in which
vT /,

collision mean free path,

24

(A.104)

(A.105)

has been defined for thermal molecules. In these closure relations, is the
heat conduction coefficient, m is the heat diffusivity and m is the viscosity
coefficient. The superscript m on the various coefficients indicate that they
arise from the micoscopic (molecular) processes of discrete collisions in the gas.
Equations (A.103) and (A.104) give the thermodynamic fluxes q, induced
by the thermodynamic forces T, W. If the appropriate Boltzmann collision
operator is used instead of the approximate Krook-type model of (A.101), the
scaling of the m , m , and m molecular diffusion coefficients with collision
frequency and thermal speed remains the same; however the numerical factors
in (A.103) and (A.104) change slightly.
The reference cross section for atomic and molecular collisions is 0
a20 1020 m2 in which a0 is the Bohr (atomic) radius (A.154). For standard
temperature and pressure (STP) air at the earths surface, the average cross
section for molecular collisions is 40 0 4 1019 m2 , the density is nn
2.5 1025 m3 , and the thermal speed is vT 300 m/s. Thus, for standard air
nvT 3109 s1 , vT / 107 m, and m /nm 2 /2 1.5105
m2 /s, m (5/2)(m /nm).
The Chapman-Enskog analysis is valid as long as the collision mean free
path is short compared to the gradient scale lengths (i.e., | ln T | << 1,
|V|/|V| << 1) and temporal variations are slow compared to the collision
time [e.g., 1 ( ln n/t) << 1]. Substituting the closure relations given in
(A.103) and (A.104) into the momentum and energy conservation equations
yields (neglecting for simplicity the small effects due to gradients of the transport
coefficients m and m ):
$
%
m
dV

= nF p
V + m 2 V,
(A.106)
mn
dt
3
m

m
3 dT
2
+ T (V) = m 2 T + 2
|W| .
(A.107)
2 dt
n
The diffusive components of these equations indicate that the molecular diffusion coefficients for momentum (viscous) and heat diffusion are m /nm and m ,
both of which scale as 2 and have units of m2 /s. A physical interpretation of
the processes and parametric scalings that underly these diffusion coefficients
are given in the next section.
An equation can also be developed for the evolution of the collisional entropy
s which is dimensionless and is defined in kinetic theory for f ' f0CE by
$ 3/2 %
+
T
1
d3v f ln f = ln
+ constant, collisional entropy. (A.108)
s
n
n
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

25

Note that this entropy is the negative of the Boltzmann HB function [see (A.85)]
and yields the same result as that obtained from equilibrium statistical mechanics [see (A.76)]. Taking the total time derivative of this equation yields, upon
substituting in (A.90) and (A.92),
!
"
3 dT
dn
(ns)
ds
= n
T
= T
+ (nsV)
nT
dt
2 dt
dt
t
= q :V,
entropy evolution.
(A.109)
Alternatively, since the flow of entropy density (entropy flux) is nsV + q/T ,
after using the density conservation relation (A.90) and the rate-of-strain tensor
definition (A.100),
(ns)/t + (nsV + q/T ) = (1/T ) [ q ln T + : W ]

(A.110)

in which represents the rate of entropy production due to dissipative (irreversible) processes, which is positive definite and caused by the thermodynamic
fluxes q, flowing in response to the thermodynamic forces ln T , W.
For the closure relations given in (A.103) and (A.104) the entropy production rate simplifies to (again neglecting gradients in the transport coefficients
m , m ):
2

= n m | ln T | + 2 m |W| ,

entropy production rate.

(A.111)

Thus, entropy is produced by the microcopic collisional processes that diffusively


relax the gradients of the temperature and flow velocity in the gas. The entropy
production rate is small under the Chapman-Enskog approximations (large ,
2
small = vT /): ds/dt 2 | ln T | << if heat conduction effects are
2
dominant, or ds/dt 2 |V| /vT2 << if viscous flow damping is dominant. Hence, for processes that are rapid compared to the collisional entropy
production rate and where the entropy flow induced by the conductive heat
flux q is negligible (e.g., in a constant temperature gas), it is sufficient to use
the adiabatic or isentropic (i.e., non-dissipative, constant entropy) equation
of state for an ideal gas obtained from setting ds/dt = 0:
$ 3/2 %
$ 3/2 %
T
d
p
p
d
ln
=
ln
= 0 =
= constant,
dt
n
dt
n
n5/2
isentropic (adiabatic) equation of state, (A.112)
where is 5/3 for the three-dimensional system being considered, but in general
is given by = (N + 2)/N in which N is the number of degrees of freedom in
the system. [In thermodynamics cP /cV is the ratio of the heat capacity
( u/T ) at constant pressure to that at constant volume.] Note that for a
constant density gas of volume V = 1/n, (A.112) becomes the familiar equation
of state for an ideal gas: pV = constant. The adiabatic or isentropic equation
of state can be used in place of the energy balance equation (A.92) or (A.107)
for studies of rapid, isentropic processes because there is no significant entropy
production or heat flow for such processes.
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

26

Most of the previously noted standard textbooks on statistical mechanics


provide intermediate level descriptions of the kinetic theory of gases. Advanced
level textbooks and monographs that deal specifically with the kinetic theory
of gases include:
Chapman and Cowling, The Mathematical Theory of Non-Uniform Gases (1952)
[?]
H. Grad, Principles of the Kinetic Theory of Gases, in Handbuch der Physik ,
Volume 12 (Springer-Verlag, Berlin, 1957) [?]
R. Herdan and B. S. Liley, Dynamical Equations and Transport Relationships
for a Thermal Plasma, Rev. Mod. Phys. 32, 731 (1960). [?]

A.5

Stochastic Processes, Diffusion

The heat and momentum diffusion produced by the collision-induced random


steps or motions of molecules in a gas can be understood in terms of a stochastic
or random walk process. Such processes are often called Brownian motion (after
a botanist Robert Brown who, in 1827, observed irregular motions of small
colloidal size particles immersed in a fluid), or more formally a Markoff process
(no memory of previous history or steps).
For a simple one-dimensional mathematical model of the random walk process, assume that between collisions (or another random process) a molecule
moves a distance x in a random direction (to the right or left) in a time t.
For such a process the position xn of a molecule after the nth step is related to
the position xn1 after the previous step by
xn = xn1 + Rn x

(A.113)

in which Rn is randomly 1. Using this mapping equation as a recursion relation, one finds that after N random steps the difference of the final position xN
from the initial position x0 becomes
xN = x0 + x

N
7

Rn .

(A.114)

n=1

In the limit of a large number N of random steps one obtains:


'
'
:
' xN x0 '
| n Rn |
O(1)
'
'
= lim
= lim
= 0,
lim
N ' N x '
N
N

N
N

(A.115)

because Rn is randomly 1. Thus, after a large number N of random steps


the average position of a molecule does not deviate much (<< N x) from its
initial position x0 .
However, as illustrated in Fig. A.2, the random steps do have an effect:
they cause such molecules to wander randomly in the x direction, to ever larger
distances from x0 as the number N of random steps increases. Thus, after a large
number of random steps the position of a molecule is described by a probability
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

27

Figure A.2: Illustration of the random walk process of molecules stepping a


distance x randomly to larger or smaller x for a total of N = [t/t] steps: a)
an example of a detailed particle trajectory; b) distribution of particle positions
over the N steps. The smooth dashed curves represent theN analytic
formulas given in (A.117) and (A.122) with = 4Dt x 2N .
distribution peaked at the initial position x0 with a spatial spread that increases
with N and a peak magnitude that decreases with N see Fig. A.2.
To quantify the spatial spreading effect, and hence the width of the probability distribution, one uses the first form of (A.114) to calculate the square of
the difference of the final from the initial spatial position:

&N
(2
N
N
N
7
7
7
7
Rn = (x)2
Rn2 +
Ri
Rn . (A.116)
(xN x0 )2 = (x)2
n=1

n=1

i=1

n'=i

In the limit of a large number N of random steps, the mean spread is given by

N
N
N
(xN x0 )2
1 7 2 7 7
N + O(1)
= 1,
= lim
Rn +
Ri
Rn = lim
lim
2
N N (x)
N N
N
N
n=1
i=1
n'=i

(A.117)

because Rn2 = (1)2 = 1. Hence, the average square of the spatial spreading
after a large number N of random steps (or a time t = N t) will be given by
(xN x0 )2 ' N (x)2 = t

(x)2
,
t

or

d (xN x0 )2
(x)2
'
.
dt
t

(A.118)

In summary, a random walk process produces a spatial spreading, which is called


stochastic diffusion or simply diffusion, of molecules about their initial position,
but no net motion of the average position of the molecules.
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

28

However, in an inhomogeneous medium there is, on average, a net motion


or flux of particles. The particle transport flux produced by a large number
of molecules undergoing such random walk processes in a neutral gas with a
spatially varying density n = n(x) can obtained as follows. In a time t the
plane x = x0 will be traversed by the half (on average) of the molecules that
experience collisions in the layer between x0 x and x0 , and which are moving
to the right (+). Thus, the flux ( nV = n dx/dt) of molecules moving to the
right is
?
'
!
+
x
1 x0
dx
x dn ''
=
n(x0 )
n(x)
+
+ =
2 x0 x
t
2t
2 dx 'x0

in which the density n(x) has been expanded in a Taylor series about x0 . Similarly, the flux of molecules moving through the x = x0 plane to the left ()
is
?
'
!
+
1 x0 +x
dx
x
x dn ''
n(x)
=
n(x0 ) +
+ .
=
2 x0
t
2t
2 dx 'x0
The net particle flux is the difference between these two fluxes:
'
dn ''
, Ficks diffusion law.
= + = D
dx 'x0

(A.119)

(x)2
,
2 t

(A.120)

For the simple model being considered D is given by


D =

diffusion coefficient,

which has units of m2 /s. Thus, the diffusion coefficient D is half the rate of
spatial spreading for a random walk process see (A.118).
The natural step size x for the motion of molecules between collisions in
a neutral gas is , the collision mean free path. The characteristic time t
between collisions of molecules is 1/. Thus, one infers from (A.120) that the
scaling of diffusivities induced by molecular collisions should be D 2 , which
was what was obtained in (A.103) and (A.104) in the preceding section. In a
monoatomic neutral gas there are heat and momentum diffusivities but there is
no particle diffusivity (or particle flux ) because, while two colliding molecules
exchange energy and momentum during the molecular collisions, the density of
molecules is usually unchanged as a result of the collisions.
In more realistic situations different molecules may have different x and
t values; then one must take an appropriate average and D = .(x)2 //(2t).
Since the parametric scaling of the diffusion coefficient is quite general, but appropriate averages are often difficult to formulate or evaluate for various physical
processes, the expression for D in (A.120) is mostly used to infer the scaling
of the diffusion coefficient with physical parameters. Then, kinetic calculations
are used to obtain the relevant numerical coefficients the headache factors.
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

29

In the presence of this random walk process induced by molecular collisions,


the equation for the density n(x, t) [see (A.90] becomes n/t + (nVx )/x = 0.
Using the Ficks diffusion law (A.116) for nVx = yields a one-dimensional
diffusion equation:
n
n
=
D ,
t
x x

diffusion equation.

(A.121)

To illustrate the properties of solutions of this equation, imagine that a small


number N of molecules are added to the medium at the position x0 : n(x, 0) =
N (xx0 ). After a short time, the appropriate (Green-function-type) solution
of (A.121) is
2

n(x, t) = N

e(xx0 ) /4Dt
,

4Dt

short-time diffusive distribution,

(A.122)

as can be verified by direct substitution. Note that this distribution of particles


has the desired properties for a random walk process and represents it well in
the N limit see Fig. A.2b. In particular, it is peaked at x = x0 and
spreads spatially and decreases in magnitude as time progresses. In a time t
(assumed >> t) the average spreading of the molecules in the x direction is
)
)
2
dx (x x0 )2 n(x, t)
dy y 2 ey
(x)2

0
2
)
(x x0 )
= 4Dt )
=
2Dt
=
t
t
dx n(x, t)
dy ey2

0
(A.123)

in which y (x x0 )/ 4Dt and the integrals have been evaluated using (??).
Note that this rate of spatial spreading of the density agrees with that inferred
above for the random walk process of a molecule (A.118).
The Gaussian character of this distribution can be emphasized by writing
the short time diffusive density response in (A.122) in the form
2

e(xx0 ) /

,
n(x, t) = N

4Dt.

(A.124)

In this form one readily sees from (??) that in the t 0 limit the solution
becomes a delta function (see Section B.2) at x = x0 : limt0 n(x, t)=N (x
x0 ), which was the initial condition. Also, in terms of the root mean square
spatial spread becomes simply
@
0
11/2

t
2
= x N ,
= 2Dt = = x
xrms (x x0 )
t
2
root mean square spatial spread. (A.125)
The last result shows that, as indicated in Fig. A.2, the spatial spreading is
proportional to the square root of the number of random walk steps.

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

30

Using these formulas, note also that the average time t required for molecules
to diffuse a distance x x x0 from the initial position x0 is
t

(x)2

2D

x
x

%2

t,

time to diffuse a distance x.

(A.126)

Hence, the time t required to diffuse a short distance x is the product of the basic random walk time t times (x/x)2 the square of the number of random
walk steps x in the distance x to be traversed. This quadratic dependence
of the spreading time on the spreading distance is an intrinsic property of diffusive processes. As a caveat on this analysis, note that the solution (A.122)
is only valid for short times: t < L2n /D (Ln /x)2 t so the background
medium density
# coefficient are reasonably constant over the disand diffusion

tance / 2 = 2Dt = x t/t that typical particles spread over in the time
t [i.e., (/n)(dn/dx) /Ln << 1 and (/D)(dD/dx) << 1].
In a finite box, as time progresses molecules eventually diffuse to the boundaries of the box where it will be assumed the molecules are absorbed. The
question then becomes: what is the average confinement time for molecules in
the box? Assume for simplicity that: the diffusion coefficient is constant in
space; a one-dimensional treatment is sufficient; N molecules are inserted at
the center (x = 0) of a box of width 2L (assumed >> x = ) at time t = 0;
and the density of molecules vanishes at the box boundaries (x = L). Then,
the solution of the diffusion equation (A.121) for this boundary value problem
can be shown (by separation of variables, expansion in sinusoidal eigenfunctions)
to be
n(x, t) =

0 x1
(2j + 1)
L2
N 7 t/j
, with j
, j 2 . (A.127)
e
cos j
L j=0
L
2
j D

For short times (t << 0 ) the box boundaries at x = L are unimportant


and this solution reduces to (A.124), which is a more convenient form then.
For intermediate times (t j ) the sinusoidal eigenfunctions (up to at least
2j) must be summed to obtain the response. In the time asymptotic limit
(t > 0 > 1 > 2 ) the lowest order eigenmode solution dominates:
n(x, t)

t>0

'

0 x 1
N t/0
e
.
cos
L
2L

(A.128)

Thus, an average confinement time for molecules in the box can be identified
as
0

L2
L2
L2
=
'
,
2
2
0 D
(/2) D
2.5D

confinement time.

(A.129)

Note that upon using D = (x)2 /2t one obtains 0 = (2/20 )(L/x)2 t,
which quantifies (for this specific case where x L, t 0 ) the headache
factors in the scaling relation (A.126). For cylindrical, spherical boxes the
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

31

eigenfunctions are Bessel, spherical Bessel functions and the lowest order eigenvalues are 0 ' 2.405, 0 = , respectively. Then, using a as the radius of the
box, one obtains confinement times of 0 ' a2 /6D, a2 /10D for cylindrical,
spherical systems, respectively.
Many of the references noted at the end of the two preceding sections have
discussions of random walk (Brownian motion) and stochastic diffusion processes. The classic and ageless reference for such processes is:
S. Chandrasekhar, Stochastic Problems in Physics and Astronomy, Rev. Mod.
Phys. 15, 1 (1943). [?]

A.6

Fluid Mechanics

The equations of hydrodynamics used to describe the behavior of a fluid are


the fluid moment equations obtained from the kinetic theory of gases (A.90),
(A.91) and (A.112). However, they are usually modified by writing them in
terms of the mass density m n m, which has units of kg/m3 :
mass continuity equation:
Navier-Stokes equation:
(momentum balance)
isentropic equation of state:

m
+ m V = 0,
(A.130)
t

dV
F
= m p# + m 2 V, (A.131)
m
dt
m
d
ln
dt

p
m

= 0,

(A.132)

in which p# p (m /3)V. (The equation of state is often called the


adiabatic equation of state in hydrodynamics.) In these equations d/dt is the
total time derivative taking account both of the direct temporal derivative and
the effects of the advection by the flow velocity V in the fluid:

+ V,
dt
t

total time derivative.

(A.133)

For gases or liquids in the earths atmosphere the relevant force on molecules
is the gravitational force, which is a conservative force:
G = FG = mVG ' mg
ex m g,
F

gravitational force,

(A.134)

where VG = ME G/R is the gravitational potential. In the last expression use


has been made of the fact that near the earths surface (x R RE << RE ,
2
)(RE x + ). Also, here,
radius of the earth) one has VG ' (ME G/RE
2
) ' 9.81 m/s2 is the gravitational acceleration at the earths
g (ME G/RE
surface.
The velocity flow field V can in general be decomposed into parts representable in terms of scalar, vector potentials , C (see Section D.5):
V = + C,
DRAFT 11:16
September 2, 2003

potential representation of a flow field.

c
#J.D
Callen,

(A.135)

Fundamentals of Plasma Physics

32

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

The scalar potential part represents the longitudinal, irrotational or compressible part of the flow since V = = 2 . The vector potential
part is incompressible since C = 0. However, this component represents
rotation or vorticity5 (units of s1 ) in the flow:
V = (C),

vorticity.

(A.136)

The properties of sound waves in a fluid can be illustrated by considering


compressible perturbations of air in the earths atmosphere. The equilibrium
pressure distribution is determined from the hydrostatic force balance equilibrium, which is the equilibrium (/t = 0) and small viscosity limit of the
Navier-Stokes equation that in the absence of equilibrium flows (V0 = 0) becomes simply:
0 = m0 VG p0 .
Assuming for simplicity that the temperature T is constant, taking x to be
the vertical distance above the earths surface, and using p = nT = m0 T /m
(Boyles law for this situation), the hydrostatic equilibrium becomes
0 = m0 g

T dm0
m dx

m0 (x) = m0 (0) emgx/T ,


law of atmospheres.

(A.137)

Thus, in equilibrium the density of air decreases with distance above the surface
of the earth on a scale length of T /mg = vT2 /2g 104 m.
To exhibit the properties of sound waves consider perturbations of the compressible air in this equilibrium:
m = m0 + m ,

V = V,

p = p0 + p,

perturbed equilibrium,

(A.138)

in which the tilde over a quantity indicates the perturbation in that quantity.
Substituting these forms into the fluid equations (A.130)(A.132) yields, upon
neglecting the effect of gravity for simplicity and linearizing the equations (i.e.,
neglecting all quantities that are quadratic or higher order in the perturbations):
m
+ V

+ m0 V
m0 = 0,
t
where
cH
S

#
#
p0 /m0 = T /m,

m0

=
p + m 2 V,
t

hydrodynamic sound speed,

m
p = cH
S

(A.139)

which is typically
about 340 m/s at the earths surface. Note that for a neutral
#
=
/2
v
gas cH
T . For an equilibrium that is approximately homogeneous
S
over the collision mean free path ( | ln m0 | << 1), a perturbed density
5 A physical example of vorticity is the circular flow of water around a drain in a bathtub
as it is being emptied.

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

33

0= 0). For such


response m exists only for compressible perturbations (V
perturbations these equations can be combined to yield
0 2
1
2 m
H
m 2

V
= 0.
m
S
t2
Considering perturbations that are localized relative to the scale length of the
2
m
equilibrium density gradient so that the cH
S and terms can be neglected,
but longer scale than the collision mean free path (typically 107 m for air at
2
4
the earths surface) so the viscosity can be neglected (i.e., cH
S /g 10 m >>
7
perturbation scale length >> 10 m), this equation becomes simply
2 m
2 2
cH
m = 0,
S
t2

sound wave equation.

(A.140)

< 0) and rarefy (V


> 0) the
Thus, density perturbations compress (V
fluid as they propagate through it adiabatically (with negligible entropy production) at the sound speed cH
S defined in (A.139).
To exhibit the properties of the most fundamental type of fluid instabilities,
the Rayleigh-Taylor (R-T) instabilities, consider perturbations of nearly incompressible liquids, in a case where a heavy liquid is above a lighter liquid and
0) perturbations the
the two fluids are immiscible. For incompressible (V
linearized continuity equation becomes
m

= V
advective response,
(A.141)
m0 ,
t
which indicates the change in local mass density caused by a perturbed flow
in the direction of the gradient in the equilibrium mass density. Combining
this advective response with the partial time derivative of the linearized NavierStokes equation yields
0
1

2V
p

.
m0 2 = V
m0 VG
t
t
Taking the curl of this equation to eliminate the perturbed pressure gradient and
hence the coupling to sound waves, and neglecting gradients in the equilibrium
>> m0 V)
and viscosity
compared to those in the perturbations (m0 V
effects (valid for perturbation scale lengths long compared to the collision mean

V:
free path ), yields for the perturbed flow vorticity
0
1

m0 2 = V
m0 VG .
t
Considering a coordinate system where x is directed vertically upward and y, z
are in a plane parallel to the earths surface, and assuming wavelike pertur = Cz
= C
=
ez , Cz
bations of the type V
ez Cz , in which C
exp(ikx it) with kx << ky is a stream function [see (??,??)], so that
=
ez ky2 Cz , yields an equation for the perturbation frequency:

ez 2 Cz '
2 ' VG ln m0 = g m .
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

(A.142)

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

34

When VG m0 < 0 (light liquid above heavy liquid since VG g is


upward), 2 is positive and two benign, oscillating waves occur. (Adding viscosity effects causes the waves to be damped.) However, when a heavy liquid is
placed over a light liquid (VG m0 = g m0 > 0) the 2 < 0 indicates
complex conjugate roots, one of which will be growing exponentially in time at
rate:
Im{} ' ( g ln m0 )1/2 ,

R-T instability growth rate

(A.143)

This is the Rayleigh-Taylor (or interchange) instability by which the interface


region between the upper heavy fluid and the lower lighter fluid develops growing
undulations that lead ultimately to interchange of the positions of the heavy and
light fluids.
The overall process of the interchange of the two fluids can be thought of
as consisting of the following steps. First, thermal fluctuations excite a modest
undulation of the boundary between the two liquids. If the heavy fluid is on top,
this spontaneous perturbation grows exponentially in time at the rate indicated
by (A.143). The undulations grow to a large amplitude where the linearization
procedure used to derive (A.142) becomes invalid. Lagrangian coordinates (i.e.,
coordinates that follow particular fluid elements as they move rather than the
usual fixed position Eulerian ones) can be used to explore the growth of the
structures into the slightly nonlinear regime. However, ultimately the vortexlike collective motions of the fluids become highly nonlinear, very contorted and
large enough to encounter adjacent vortices and/or the boundaries of the regions
occupied by the fluids. Then, turbulence in the fluid develops and it cascades
the large vortices into smaller ones, turbulently mixing the two fluids until the
heavier one is on the bottom.
In order to describe the behavior of the vortices as they evolve nonlinearly
towardA the turbulent state, consider the total time derivative of the circulation
CK C d" V in the rotational part of the flow V, which is responsible for the
vortex, over the closed curve C within the fluid:
dCK
dt

*
++
d
d

d" V =
dS V
dt
dt S
++
++ C
dS [(/t + V) (V)] +
(dS/dt) V
=
S
S
++
dS {(V/t) [V (V)]}
=
++S
dS [/t (V)]
(A.144)
=
S

in which use has been made of Stokes theorem (??), S is the open surface
bounded by the closed curve C that moves with the encompassed fluid, dS/dt =
(V)dS V dS [see (??)], and the vector identity in (??) has been used.
Dividing the Navier-Stokes equation (A.131) with a conservative force F by m
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

35

and taking its curl to obtain an equation for V/t yields, after making
use of (??):
$
!
"
%
++
m 2
1
m
dCK
=
(V) +
dS
m p
V . (A.145)
dt
2m
3
m
S
For an adiabatic equation of state (A.132), p m and hence m p = 0.
Thus, the circulation CK is constant in time, except for the dissipative effects
due to viscosity that are small for all but very short scale lengths of the order of
the collision mean free path because m /m 2 . Thus, on most relevant
scale lengths
dCK
= 0 for m 0,
dt

Kelvins circulation theorem,

(A.146)

for inviscid (zero viscosity) fluids.


What this theorem shows is that a vortex tube moves with (or is frozen
into) the fluid as it evolves, and that the amount of circulation CK in the flow
field V remains constant except for the effects of viscosity, which becomes
important in boundary layers near the edge of the fluid or at the edge of vortices
that come close to other vortices. However, the derivation relied on the use of
Stokess theorem, which required that the topology of the closed curve C be
continuous and that it remain so. Thus, the invariance of CK could be broken
by nonlinear interactions between vortex structures that break or reconnect the
topology by causing the bounding curve C, which is ))
expected to always move
with the fluid and encompass the same vorticity flux S dS V, to become
discontinuous. To the extent that the topology of the surfaces of vorticity flux
remains intact there is no motion (or transport) of fluid relative to these surfaces.
However, the flux surfaces of the vorticity can distort in shape as they move
around in the fluid. Thus, vortex tubes or eddies are relatively robust objects
in low viscosity fluids.
The nonlinear evolution and interactions of vortices in a fluid are governed
by the vorticity evolution equation
m 2

= (V) (V) +

t
m
or,
d

dt

m 2
+ V = V +
.
t
m

(A.147)

(A.148)

These equations are obtained by taking the curl of (A.131), which eliminates the
coupling to sound waves, and assuming for simplicity that the mass density is
constant. The (V) term in (A.147) represents the advection of the vorticity vector by the flow velocity V as indicated by the last line of (A.144).
The V term on the right of (A.148) represents vortex tube stretching by
gradients in the velocity flow; it vanishes for two-dimensional flows. In threedimensional flows the vortex tube stretching term reduces the area of a vortex
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

36

but also increases its vorticity to keep the vorticity flux CK constant as
required by Kelvins circulation theorem.
The ratio of the nonlinear advection of vorticity [(V), a nonlinear
inertia term] to the viscous dissipation of vorticity (m 2 ) is
Re =

m0 V0 L0
(V)

,
m
2
( /m )
m

Reynolds number,

(A.149)

in which V0 , L0 are typical flow speeds and gradient scale lengths in the fluid.
When the vorticity evolution equation is written in terms of dimensionless variables, the reciprocal of the Reynolds number is the only dimensionless parameter
in the equation as the coefficient of the viscous dissipation term. For example,
for incompressible flows (i.e., ones that do not excite sound waves and are dominated by vorticies), Eq. (A.147) can be written in terms of the dimensionless
V/V0 , and
L0 as
variables t (V0 /L0 )t, V
1 2

= (
V)
+
.
t
Re

(A.150)

Thus, all incompressible flows with the same Reynolds number and the same flow
< 1)
geometry will have the same flow properties. At low Reynolds numbers (Re
the flow is laminar. For not too large Reynolds numbers vortex structures
induced by the particular geometrical situation (e.g., flow past a fixed body)
> 103 ) the
tend to dominate the flow pattern. For high Reynolds numbers (Re
nonlinear vorticity advection overwhelms the viscous dissipation and the flow
becomes turbulent.
In fully developed turbulence (Re >> 103 ) there is a cascade of energy
from macroscopically-induced large-scale vortices through nonlinear interactions
of turbulent eddies of successively smaller dimensions until the scale lengths
become so small that the energy in the eddies is viscously dissipated. (The
effective Reynolds number is close to unity for the dissipative scale eddies.)
Since the dominant eddy interaction term is the vortex stretching term V in
(A.148), successive generations of the turbulent eddies become longer, thinner
and have larger vorticities. Thus, the mean square vorticity, which is known as
the enstrophy ( ||2 ), increases during the cascade.
For sufficiently large Reynolds numbers there is a large inertial range of
spatial scale lengths for which the vortex interactions are predominantly nonlinear (i.e., where the large-scale stirring and small-scale viscous dissipation
effects are negligible). In the inertial range the turbulent eddies are self-similar
(i.e., of the same structure, independent of scale size, from one generation to the
next one). The energy flow per unit mass = (V)(V 2 /2) from one wavenumber range k to the next smaller one can be estimated by k kVk3 Vk2 /k
in which k (kVk )1 is the turbulent decorrelation or eddy turnover time
at a given k. Since energy is input via stirring at large scales and dissipated at small scales, in steady-state the energy transfer rate from one scale
to the next smaller one in the inertial range must be nearly constant. Thus,
k ' V03 /L0 , a constant for a given externally driven situation, and hence
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

37

Vk ( /k)1/3 . The energy E(k) in the turbulent fluctuations between k and


)
2/3
2k is given approximately by dk E(k) k E(k) Vk2 or E(k) k 5/3 ,
which is the Kolomogorov spectrum for turbulence within a large inertial range.
The successively smaller scale eddies have smaller velocities and energies, but
larger vorticity and faster turnover rates [k (kVk )1 ( k 2 )1/3 ] to keep
the energy flow rate in k-space constant.
In the inertial range the turbulent eddies lose their momentum on a mixing
length scale k Vk k 1/k. This leads to a Prandtl mixing length estimate for the effective diffusion coefficient [cf., (A.120)] for turbulent viscosity
1/4
in the fluid of Deff eff /m 2k /k /k 4/3 . However, this turbulent mixing is actually dissipationless; all it does is transfer the momentum
and energy to shorter scale lengths. Eventually, the eddies reach the (Kolomogorov) dissipation scale kd1 at which 1/kd (m /m ) kd2 , which yields
1/4
kd /(m /m )3/4 (Re)3/4 /L0 .
Because the effects of viscosity are negligible in the inertial range and because
the viscous dissipation scale length is so short [kd1 L0 /(Re)3/4 << L0 ], it
is tempting to neglect it entirely. However, while its effects can be neglected
for inertial range scale lengths (1/L0 << k << kd ), it must be retained in
general because it: 1) increases the order of the differential equation governing
vorticity; 2) is important in boundary layers near material objects and other
nearby vortices; and, 3) most importantly for computer simulation, provides
the only energy sink (at high k) for turbulent fluctuations in a neutral fluid.
Most of the previously noted standard textbooks on mechanics, statistical
mechanics and kinetic theory of gases contain introductory or intermediate level
descriptions of fluid mechanics. Advanced level monographs and textbooks that
specifically deal with fluid mechanics include:
Batchelor, Introduction to Fluid Dynamics (1967) [?]
Tennekus and Lumley, A First Course in Turbulence (1972) [?].

A.7

Quantum Mechanical Effects

The fundamental concept in quantum mechanics is that, owing to the wavelike


nature of particles on small scale lengths, a particles position q and canonically
conjugate momentum p cannot simultaneously be known to arbitrarily high accuracy. Rather, the product of the uncertainties in the position and momentum,
q and p, respectively, must be Plancks constant or greater:
p q h,

Heisenberg uncertainty principle.

(A.151)

This relation shows the limit of applicability of mechanical causality. The uncertainty principle holds for any pair of canonically conjugate variables. Thus,
it applies for energy and time, which for conservative systems are canonically
conjugate variables (p = H = and q = t), as well:
t h.
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

(A.152)

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

38

By Heisenbergs uncertainty principle, the position of a nonrelativistic particle


moving with velocity v in a force-free region (so that its canonical momentum
p is simply mv) cannot be known to within
h = h/mv,

de Broglie wavelength.

(A.153)

In the Bohr model of the hydrogen atom an electron gyrates at constant


radius a0 around the proton nucleus of the atom. Since the rotational angle is
a symmetry coordinate and hence totally uncertain, the Heisenberg uncertainty
principle requires that the canonically conjugate action J be quantized to integer
multiples (n) of Plancks constant:
*
*
J =
p dq =
p d = 2 me a20 0 = n h
in which the angular momentum p me a20 0 where 0 d/dt is the constant rotation frequency. The equilibrium radial force balance between the
electric field force e2 /({4+0 }a20 ) and the centripital acceleration force me a0 02
on the electron yields the equation
e2
= me a0 02 .
{4+0 }a20
Solving these two simultaneous equations for a0 in the ground state (n = 1)
case yields the characteristic radius of the hydrogen atom:
a0 = {4+0 }

(h/2)2
' 0.5291010 m,
me e2

Bohr radius.

(A.154)

This is the characteristic scale length for the size of all atoms the range
over which their electrostatic force field extends. The corresponding range over
which nuclear forces extend is
re =

e2
' 2.821015 m,
{4+0 }me c2

classical electron radius,

(A.155)

which is inferred from equating the electric potential energy e2 /({4+0 }re )
from a distributed electron charge to the electron rest mass energy me c2 .
The binding energy of an electron in a Bohr atom in its ground (lowest
energy) state is given by the (negative of the) potential energy of the electron
when it is located at the Bohr radius from the proton plus the kinetic energy of
the electron:
H
=
E

e2
me 2 2
1
me e4
a0 0 =

' 13.6 eV,


{4+0 }a0
2
{4+0 }2 2(h/2)2
Bohr atom binding energy,

(A.156)

which is also called the Rydberg energy. For electrons in the nth quantum
state the orbit radius increases by a factor of n2 and the rotation frequency
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

39

0 decreases by a factor of 1/n3 ; consequently, the binding energy of the state


decreases by a factor of 1/n2 . For electrons gyrating around an ion of charge Zi ,
the potential and consequently the electric field force increases by a factor of Zi .
This causes the Bohr radius to decrease by a factor of 1/Zi and the ionization
energy to increase by a factor of Zi2 . Thus, neglecting fine-structure effects, the
binding energy of an outer electron in a level labeled by the quantum number
n ( 1) which is gyrating around an ion of charge Zi is given by
Z
H
(n) ' Zi2 E
/n2 ,
E

outer electron binding energy.

(A.157)

Note that for a nucleus with a high atomic number Z the binding energy of
the most tightly bound (n = 1, ground state) electron, which is the last one
to be removed as an atom is ionized, can be very large. For example, for iron
(Z = 26) the binding energy of the last electron is 9 keV while for tungsten
(Z = 74) it is 75 keV.
The degree of ionization in a plasma can be estimated from the Saha equation
which gives the population density of a particular ionization and quantum state
of an atom in a gas in thermodynamic equilibrium. It can be obtained by equating the rates of ionization [ nn exp(Ui /Te )] and recombination [ ni (ne 3h )]
for ions in a partially ionized gas:
$
%3/2
2 2me Te
ne 3De
ni
'
eUi /Te = 25/2
eUi /Te
2
nn
ne
h
(ne a30 )1/2
6 1027
3/2
[Te (eV)] eUi /Te ,
'
Saha equation,
ne (m3 )

(A.158)

in which ne , ni and nn are the electron, ion and neutral density, respectively,
Ui is the ionization potential and Te is the temperature in electron volts of the
assumed Maxwellian distribution of electrons. The ionization potential Ui for
ionization of an atom from its ground (neutral) state to the first ionized state
is given by the electron binding energy in the atom [cf., Eq. (A.157)]. It ranges
from 3.9 eV for Cesium atoms to 24.6 eV for Helium.
The fractional ionization [ ni /(nn + ni )] is exponentially small for electron
temperatures Te much lower than the ionization potential Ui . The electron
temperature required to attain a small degree of ionization ( ni /nn << 1) can
be estimated by solving the Saha equation iteratively for Te :
Te |ion '

ln

Ui
% (0.021) Ui ,
6 1027 [Te (eV)]3/2
(ni /nn ) ne (m3 )

(A.159)

where in the last form the smallest number correponds to interplanetary densities ( 106 m3 ) and the largest to solid densities ( 1029 m3 ). The Te required
to produce a fully ionized state (ni /nn >> 1) is not much larger. Thus, for example, a nitrogen gas (Ui = 14.5 eV) at a density of 2.5 1025 m3 (the density
of room temperature air) becomes 1% ionized at Te ' 1.4 eV, and fully ionized
> 2.2 eV. At lower densities the electron temperature range over which
for Te
DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

40

the transition from a partially to fully ionized gas takes place is even narrower.
For some examples of the variation with electron density of the Te required for
complete ionization, see Fig. ?? at the end of Chapter 1.
Note however that the ions might not be fully stripped of their electrons. In
particular, for Te 0.110 keV, high Z ions might not be fully stripped because
of the very large binding energy of their most tightly bound electrons. Such
ions would have an ion charge Zi < Z.
Some standard introductory level quantum mechanics textbooks are:
Krane, Modern Physics ( ) [?]
Sproul and Phillips, Modern Physics ( ) [?]
Tipler, Modern Physics ( ) [?]
Gasiorowicz, Quantum Physics ( ) [?]
Powell and Crasemann, Quantum Mechanics ( ) [?].

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

A.8

41

Physical Constants
Fundamental Physical Constants

Relative
Uncertainty
(106 )

Quantity

Symbol

Best Value6

electron mass
proton mass
elementary charge
speed of light in vacuum
permeability of vacuum
permittivity of vacuum
gravitational constant
Planck constant
Boltzmann constant

me
mp
e
c
0
+0
G
h
kB

9.109 389 7 1031 kg


1.672 623 1 1027 kg
1.602 177 33 1019 C
299 792 458 m/s
4 107 H/m
1/0 c2 F/m
6.672 59 1011 N m2 /kg2
6.626 075 5 1034 J s
1.380 658 1023 J/K

0.59
0.59
0.30
exact7
exact
exact
128
0.60
8.5

SI Units And Their Abbreviations, Interrelationships


Quantity

Name

length
mass
time
electric current
temperature
amount of substance
atomic unit of energy
atomic unit of mass
frequency
force
pressure, stress
energy, work
power
electric charge
electric potential
capacitance
electrical resistance
magnetic flux
magnetic flux density8
inductance

meter
kilogram
second
ampere
kelvin
mole
electron volt
amu
hertz
newton
pascal
joule
watt
coulomb
volt
farad
ohm
weber
tesla
henry

Symbol
m
kg
s
A
K
mol
eV
u
Hz
N
Pa
J
W
C
V
F

Wb
T
H

In Terms Of Other Units


102 cm = 1010
A
103 g
C/s
' 1/11 604.4 eV
' 1.602 177 33 1019 J
' 1.660 540 2 1027 kg
s1 (cycles per second)
m kg s2
N/m2 = m1 kg s2
N m = m2 kg s2
J/s = m2 kg s3
sA
W/A = m2 kg s3 A1
C/V = m2 kg1 s4 A2
V/A = m2 kg s3 A2
V s = m2 kg s2 A1
Wb/m2 = kg s2 A1
Wb/A = m2 kg s2 A2

6 E.R.

Cohen and B.N. Taylor, Physics Today, August 1998, BG7 [?].
speed of light fixes the length of the meter in terms of the second.
8 In plasma physics magnetic field strengths are often quoted in Gauss: 1 Tesla 10 kGauss.
7 The

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX A. PHYSICS EQUATIONS, UNITS, AND CONSTANTS

42

Other Physical Constants


Quantity
Avogadro constant
Molar gas constant
Air (20o C and 1 atmosphere)
density
sound speed
atmospheric pressure
molecular weight
viscous diffusivity
Water
density
sound speed
viscous diffusivity
Earth
mass
mean radius
gravitational acceleration
magnetic dipole moment

Symbol

Value

NA
R

6.022 1023 #/mol


8.31 J mol1 K1

n
cS
p

2.49 1025 molecules/m3


343 m/s
760 Torr = 1.01 105 Pa
28.9 g/mol
1.5 105 m2 /s

m /m
n
cS
m /m

3.33 1028 molecules/m3


1460 m/s
106 m2 /s

ME
RE
g
Md

5.98 1024 kg
6.37 106 m
9.81 m/s2
8.0 1022 A m2

Particle Masses
Particle
or Atom

Symbol

electron
muon
proton
neutron
deuteron
triton
helium
carbon
nitrogen
oxygen
argon
iron
molybdenum
tungsten

me
m
mp
mn
mD
mT
mHe
mC
mN
mO
mAr
mF e
mM o
mW

Atomic
Number Z

1
1
1
2
6
7
8
18
26
42
74

Best Atomic
Mass9 Value
0.000 548 579 903
0.113 428 913
1.007 276 470
1.008 664 904
2.013 553 214
3.016 050
4.002 603
12.011 15
14.006 7
15.999 4
39.948
55.845
95.94
183.84

Energy Units
(mc2 /e, MeV)
0.511
105.658
938.272
939.566
1 875.613
2 809.853
3 728.402

9 The unified atomic mass unit = 1.660 540 2 1027 kg (0.59 106 relative error) =
931.494 32 MeV (0.30 107 relative error). Note also that Avogadros constant NA 1/u.

DRAFT 11:16
September 2, 2003

c
#J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX B. SPECIAL FUNCTIONS

Appendix B

Special Functions
Key properties of special functions as they are often used in plasma physics are
summarized in this appendix.

B.1

Heaviside Step Function

The Heaviside step function is usually just called the step function. It was introduced by Oliver Heaviside in the late 1800s to represent the idealized switching
on (in time) of a voltage or other source in electrical engineering problems. It
is defined by
!
1, x > x0 ,
(B.1)
H(x x0 ) =
0, x < x0 .

At x = x0 , H is in general undefined (because it is discontinuous there), but


will be taken to be 1/2 there so that it is equal to its average value at this jump
discontinuity. The derivative of the Heaviside step function is the Dirac delta
function:
(B.2)
H ! (x x0 ) = (x x0 )
in which the prime denotes differentiation with respect to the argument. The x
dependence of the step and delta functions are shown schematically in Fig. B.1.

B.2

Dirac Delta Function

The Dirac delta function, which is usually just called the delta function, is a
concentrated spike or impulse of unit area. It was introduced by P.A.M. Dirac
in the 1920s in the context of developing a physical interpretation of quantum
mechanics. The delta function is often used in plasma physics to represent the
spatial distribution of point charged particles. It also arises naturally from
integrals that yield singular responses to resonant perturbations.

DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX B. SPECIAL FUNCTIONS

Figure B.1: Schematic of x dependence of the Heaviside step function and Dirac
delta function.

The one-dimensional delta function is defined by the following properties:


(x x0 ) = 0 for x #= x0 ,
!
" b
f (x0 ), a < x0 < b,
dx f (x) (x x0 ) =
0,
otherwise,
a

(B.3)

for any function f (x) that is continuous at x = x0 . Thus, the delta function
is zero except at the point where its argument vanishes; there it is so large
(i.e., singular) that the integral of it over that point (its area) is unity [i.e., for
#b
f (x) = 1 we have a dx (x x0 ) = 1]. Note that hence the delta function has
units of one over the units of its argument.
The delta function is a mathematically improper function because it is
unbounded where its argument vanishes. However, it is a generalized function
whose integral can be defined through a limiting process in distribution theory.
Specifically, for a unity area distribution function w(x), one defines a delta
sequence w(x; ) which becomes progressively more peaked (height 1/)
and narrower (width ) in the limit that 0 such that it becomes a unit
area spike. In terms of such a delta sequence, one defines
"

dx (x) f (x) lim

"

dx w(x; ) f (x) = f (0).

(B.4)

Symbolically, we can write


.
(x) = lim w(x; ),
0

(B.5)

which is only valid in evaluating integrals in the form given in (B.4).


A function w(x) is
a suitable basis for a delta sequence if it is nonnegative
#
and has unity area [ dx w(x) = 1]. Oscillatory, unity area functions that
decay as their argument increases [lim|x| |w(x)| 0] are also suitable basis
functions for delta sequences. For a suitable distribution function w(x), a delta
sequence is defined by w(x; ) w(x/)/. Examples of delta sequences based
2
on Gaussian [wG = ex / ], Lorentzian [wL = 1/(x2 + 1)] and correlation

DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX B. SPECIAL FUNCTIONS


function [wC = sin x/x] distributions are, respectively,
2
1
e(x/) ,

,
wL (x; ) =
2
(x + 2 )
sin(x/)
.
wC (x; ) =
x

wG (x; ) =

(B.6)
(B.7)
(B.8)

Basic properties of delta functions include


(B.9)
(x0 x) = (x x0 ),
(B.10)
f (x) (x x0 ) = f (x0 ) (x x0 ), x (x) = 0,
" b
dx (x x0 ) (x x1 ) = (x0 x1 ) for a < x0 , x1 < b, (B.11)
a

1
[(x x0 ) + (x + x0 )] ,
2 |x0 |
$ (x xi )
1
(a).
, (ax) =
[f (x)] =
|df
/dx|
|a|
xi
i

(x2 x20 ) =

and

(B.12)
(B.13)

in which xi are the (assumed) simple zeros of f [i.e., f (xi ) = 0]. The derivative
of a delta function is a couple, which is a positive spike followed by a negative
spike. It can be integrated by parts to yield
" b
dx f (x) ! (x x0 ) = f ! (x0 ),
(B.14)
a

where the prime denotes differentiation with respect to the argument. The effect
of the j th derivative of a delta function can be calculated by integrating by parts
j times:
" b
dx f (x) (j) (x x0 ) = (1)j f (j) (x0 ).
(B.15)
a

Differentiation properties of delta functions are

(x x0 ) =
(x x0 ),
x
x0
d[x] dx(t)
dx(t)
d
[x(t)] =
!
.
dt
dx
dt
dt

(B.16)
(B.17)

For more than one dimension one simply takes products of delta functions
in the various directions. Thus, a delta function at the point x0 (x0 , y0 , z0 )
in three-dimensional Cartesian coordinate space is written as
(x x0 ) (x x0 ) (y y0 ) (z z0 ),

Cartesian.

(B.18)

In other coordinate systems the three-dimensional delta function is just the


product of the delta functions in the new coordinates divided by the Jacobian
DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX B. SPECIAL FUNCTIONS

of the coordinate transformation. Thus, three-dimensional delta functions in


cylindrical and spherical coordinates are given, respectively, by
(r r0 ) ( 0 ) (z z0 )
, cylindrical,
r
(r r0 ) ( 0 ) ( 0 )
, spherical.
(x x0 ) =
r2 | sin |
(x x0 ) =

(B.19)
(B.20)

An integral of a three-dimensional delta function over a volume V vanishes


unless V includes the point x0 :
%
" V +V
f (x0 ) if V contains x = x0 ,
3
d x f (x) (x x0 ) =
(B.21)
0
otherwise.
V
Some key summations, integrals and differentials that result in delta functions are

einx

2 (x),

(B.22)

dk eikx

2 (x),

(B.23)

d3k eikx

(2)3 (x),

(B.24)

1
|x x0 |

= 4 (x x0 ).

(B.25)

n=
"

"

Delta functions are treated simply but rigorously in


Lighthill, Intoduction to Fourier Analysis and Generalized Functions (1958) [?]
Dennery and Krzywicki, Mathematics for Physicists, Section III.13 (1967) [?].

A comprehensive treatment of generalized functions is given in


Gelfand and Shilov, Generalized Functions, Vol. I (1964) [?].

B.3

Plasma Dispersion Function

The plasma dispersion function is defined by


1
Z(w)

"

du

eu
,
uw

Im(w) > 0.

(B.26)

Analytic continuation to Im(w) 0, which is obtained by deforming the integration contour to always pass beneath the pole at u = w (see Fig. ??), yields

DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX B. SPECIAL FUNCTIONS

Figure B.2: Behavior of the plasma dispersion function Z(w) and its derivative
Z ! (w) as a function of wR Re(w). In the figures on the left the wR dependences of the real (subscript R) and imaginary (subscript I) parts of Z and Z !
are shown for Im(w) = 0 by solid and dashed lines, respectively. Corresponding
polar plots are shown on the right, which also indicate the behavior for selected
values of Im(w) wI .

the complete specification


"
2
eu

,
Im(w) > 0,
du

uw

2
"

2
eu
Z(w)
du
+ i ew , Im(w) = 0,
P

uw

"
2

2
eu

+ 2i ew , Im(w) < 0.
du
u

(B.27)

Here, P is the Cauchy principal value operator (??) that defines (i.e., makes
convergent) the integration over the singularity at u = w when w is real.
While the definition of Z(w) might appear to be discontinuous at Im(w)
= 0, it is in fact continuous there. Its continuity there can be verified by
taking the Im(w) 0 limit [for finite Re(w) #= 0] of the forms given above
for Im(w) > 0 and Im(w) < 0, and showing that they are identical to the
Im(w) = 0 definition see (??). The behavior of the real and imaginary parts
of Z(w) and its derivative Z ! (w) are shown in Fig. B.3. As indicated in Fig. B.3,
the plasma dispersion function has the following symmetry properties:

2
(B.28)
Z(w) = 2i ew Z(w),
w2

,
(B.29)
Z(w ) = [Z(w)] = Z (w) + 2i e
where the superscript indicates the complex conjugate.
which is defined by the integral in (B.26) but
A complementary function Z,
for Im(w) < 0, is related to Z(w) by

(B.30)
Z(w)
= Z(w) = Z(w) 2i ew .
The wI Im(w) 0 representations of the plasma dispersion function and its
complement can be obtained directly using the Plemelj formulas (??).
DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX B. SPECIAL FUNCTIONS

An alternative definition of Z(w) that is valid for all finite Im(w) is


2 "
2
2
2
2 ew
Z(w)
dt et = ew [1 erf(iw)] ew erfc(iw),
W (w) = =
i

iw
(B.31)
which indicates the close relationship to the error function. This is the form
of the plasma dispersion function most commonly used in the Russian plasma
physics literature, and in error function reference books.
The plasma dispersion function satisfies the differential equation
Z ! (w)

dZ
= 2 [ 1 + w Z(w) ].
dw

(B.32)

This differential equation can be used to write higher order derivatives in terms
of lower order derivatives:
Z (n)

dn Z
dn1 (wZ)
= 2
= 2[(n 1)Z (n2) + wZ (n1) ]
n
dw
dwn1

for n 2.

(B.33)
The plasma dispersion function has a power series expansion about w = 0
given by
Z(w)

2
= i ew w

(w2 )n
(n + 1/2)
n=0

*
+
w2
4w4
8w6
2w2
+

+ .
2w 1
= i e
3
15
105

(B.34)

Its asymptotic expansion for |w| >> 1 is

where

2
1 $ (2n+1)
w
(n 1/2)
Z(w) i ew
n=0
*
+
w2
1
1
3
15
1+

+
+
+
= i e
w
2w2
4w4
8w6

0, Im(w) > 0,
1, Im(w) = 0,

2, Im(w) < 0.

(B.35)

(B.36)

Corresponding power series and asymptotic expansions for the derivative of the
plasma dispersion function are given, respectively, by
,

2
for |w| << 1 and (B.37)
Z ! = 2iw ew 2 1 2w2 +
*
+

2
1
3
+

for |w| >> 1. (B.38)


Z ! 2iw ew + 2 1 +
w
2w2
Related, but more complicated integrals of the form
"
2
1
un eu
, n 0,
du
Zn (w)
uw

DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

(B.39)

Fundamentals of Plasma Physics

APPENDIX B. SPECIAL FUNCTIONS

can be calculated in terms of derivatives of Z(w) as follows. First, taking n


successive derivatives of Z(w) and integrating by parts n times one obtains
"
"
2
2
dn Z(w)
n!
eu
1
du dn (eu )
(n)
=
du
=
.
Z (w)
dwn
(u w)n+1
dun

u w

From the Rodrigues formula for Hermite polynomials Hn (u),


2
dn u2
(e
) = (1)n eu Hn (u),
n
du

it is clear that the nth derivative of Z(w) can be written as


"
2
dn Z(w)
(1)n
Hn (u) eu

.
=
du
dwn
uw

(B.40)

Now, any power of the variable u can be expressed in terms of a series of Hermite
polynomials with orders up to and including the power of u through the relation
M
1 $
dm (n) Hn2m (u),
u = n
2 m=0
n

(B.41)

in which dm (n) are the coefficients given in Table 22.12 of Abromowitz and
Stegun [?] and the upper limit of the sum is M [n/2], the largest integer less
than or equal to n/2. Substituting (B.41) into (B.39) and utilizing (B.40) yields
Zn (w) =

[n/2]
1 $
dn2m Z(w)
n2m
(1)
d
(n)
.
m
2n m=0
dwn2m

The first four of these functions are

Z0 (w) = Z
Z1 (w) = (1/2)Z ! = 1 + wZ,
Z2 (w) = (1/4)[2Z + Z !! ] = (w/2)Z ! = w + w2 Z,
Z3 (w) = (1/8)[6Z ! + Z !!! ] = (1/2)[1 + 2w2 (1 + wZ)],

(B.42)
(B.43)
(B.44)
(B.45)

in which the primes denote differentiation with respect to the argument and in
the last equalities we have made use of the definitions of differentials of Z given
in (B.32) and (B.33).
The plasma dispersion function is tabulated in
Fried and Conte, The Plasma Dispersion Function (1961) [?].

B.4

Bessel Functions

Bessel functions arise naturally from second order differential equations in the
radial coordinate for a cylindrical geometry. Their governing differential equation is
d2 y
dy
+ (z 2 n2 ) y = 0
(B.46)
z2 2 + z
dz
dz
DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX B. SPECIAL FUNCTIONS

Figure B.3: Variation of the basic and modified Bessel functions with their
arguments.

in which z is the independent (radial coordinate) variable and n is an integer.


The basic (fundamental) solution of this differential equation is y = Jn (z), which
is the Bessel function of the first kind of order n. The linearly independent
solution Yn (z) of (B.46) is called the Bessel function of the second kind of
order n. It is singular at z = 0. Hence it is not a valid solution for most
cylindrical geometry problems where the response must be finite at the origin of
the cylindrical geometry (z r = 0). The z dependence of Jn for n = 0, 1, 2 is
shown in Fig. B.4. The first few zeros [Jn (jn,i ) 0] of the fundamental Bessel
functions are (for i = 1, 2, 3, )
j0,i = 2.405, 5.520, 8.654, ,

j1,i = 3.832, 7.016, 10.173, .

(B.47)

Changing z from a real to an imaginary argument (z iz) in (B.46) changes


the sign of the z 2 term in the defining differential equation. The corresponding
solutions of this modified differential equation are the modified Bessel functions
of the first and second kind of order n, respectively: In (z) and Kn (z). The
variation of ez In (z) with the argument z is shown for n = 0, 1, 2 in Fig. B.4.
Useful recursion relations for the basic Bessel functions include
2n
Jn (z),
z
dJn (z)
,
Jn1 (z) Jn+1 (z) = 2 Jn! 2
dz
n
Jn (z) = (1) Jn (z).
Jn1 (z) + Jn+1 (z) =

(B.48)
(B.49)
(B.50)

Analogous recursion relations for the modified Bessel functions are:


2n
In (z),
z
dIn (z)
,
In1 (z) + In+1 (z) = 2 In! 2
dz
In (z) = In (z).
In1 (z) In+1 (z) =

(B.51)
(B.52)
(B.53)

A fundamental (generating function) identity that is useful for calculating


the effects on plane waves of the gyromotion of charged particles about a magDRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX B. SPECIAL FUNCTIONS


netic field is
eiz sin =

Jn (z) ein .

(B.54)

n=

The product of this equation with its complex conjugate yields


$
Jm (z) Jn (z) ei(mn) .
1=

(B.55)

m,n

Integrating this equation over from 0 to 2 yields the summation relation


1=

Jn2 (z).

(B.56)

n=

Multiplying (B.55) by cos and then integrating over and using (B.48) gives
0=

n Jn2 (z).

(B.57)

n=

By similar means it can be shown that


$
$
$
$
(1)n n Jn2 =
Jn Jn! =
n Jn Jn! =
(1)n n Jn Jn! ,
0=
n

(B.58)

1 $ ! 2
=
(Jn ) ,
2
n

$
z2
=
n2 Jn2 .
2
n

(B.59)
(B.60)

The fundamental integration identity that is useful in calculating velocityspace integrals over the product of two Bessel functions times a Maxwellian
speed distribution is
* 2
+ *
+
"
2 2
1
a + b2
ab
dx x ep x Jn (ax)Jn (bx) = 2 exp
I
(B.61)
n
2p
4p2
2p2
0
which, for a = b = s and p = 1, becomes simply
* 2+
"
2
2
1
s
dx x ex Jn2 (sx) = es /2 In
.
2
2
0

(B.62)

Summing this equation, and n times it, over n from to utilizing (B.56)
and (B.57) yields
* 2+

$
s
In
,
(B.63)
1=
2
n=
* 2+

$
s
.
(B.64)
n In
0=
2
n=
DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

10

APPENDIX B. SPECIAL FUNCTIONS

Integrals of Bessel functions with higher powers of the integration variable in


the integrand can be calculated by differentiating (B.61) with respect to p2 . For
example, with a = b = s we obtain
"
"
2 2

3 x2 2
dx x e
Jn (sx) = lim
dx x ep x Jn2 (sx)
2 1 p2
p
0
+ * 2 + * 2 + * 2 +/
.* 0
s
1 s2 /2
s2
s
s
In
+
In!
. (B.65)
= e
1
2
2
2
2
2
Power series representations (rapidly convergent for z << n) of the Bessel
functions are

m
2m
0 z 1n+2
0 z 1n $
1 0 z 1n
(1) (z/2)
1
=

+ ,
Jn (z) =
2 m=0 m! (m + n)!
n! 2
1!(n + 1)! 2
(B.66)

0 z 1n+2
0 z 1n $
1 0 z 1n
(z/2)2m
1
=
+
+ . (B.67)
In (z) =
2 m=0 m!(m + n)!
n! 2
1!(n + 1)! 2

Specific power series expansions of particular interest in plasma physics include:


z2
z4
+
,
4
64
z
z3
+ ,
J1 (z) = J0! (z) =
2 16
3z 2
,
ez I0 (z) = 1 z +
4
z2
5z 3
z
+
.
ez I1 (z) =
2
2
16

J0 (z) = 1

are

(B.68)
(B.69)
(B.70)
(B.71)

Asymptotic expansions for large arguments compared to the order (|z| >> n)
2

!
* +3
0
n 1
1
cos z

+O
,
Jn (z)
2
4
|z|
!
*
+
3
1
4n2 1
1
1
+O
ez In (z)
.
8z
|z|2
2z
2
z

(B.72)
(B.73)

The classic, comprehensive reference for Bessel functions is:


Watson, A Treatise on the Theory of Bessel Functions (1962), 2nd Edition [?]

B.5

Legendre Polynomials

Legendre polynomials are the natural (orthogonal basis) polynomials in which


to expand the latitude angle () part of a distribution function in spherical

DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

11

APPENDIX B. SPECIAL FUNCTIONS

velocity space see (??). Legendre polynomials are governed by the differential
equation ( cos )
.
/
d
2 dPl ()
(1 )
+ l(l + 1) Pl () = 0
(B.74)
d
d
and satisfy the symmetry and boundary conditions
Pl () = (1)l Pl (),

Pl (1) = 1.

(B.75)

Legendre polynomials are given in general by


Pl () =

M
1 $ (1)m (2l 2m)!
1 dl 2
l2m = l
( 1)l
l
2 m=0 m! (l m)! (l 2m)!
2 l! d l

(B.76)

in which the upper limit of the sum is M [l/2], the largest integer less than
or equal to l/2. The lowest order Legendre polynomials are
P0 = 1,

P1 = ,

P2 = (3 2 1)/2,

P3 = (5 3 3)/2.

(B.77)

Useful recurrence relations are


(l + 1)Pl+1 () + l Pl1 () = (2l + 1) Pl (),
l Pl () l Pl1 () = ( 2 1)

dPl ()
.
d

(B.78)
(B.79)

The orthgonality and values of relevant angular integrals of products of Legendre


polynomials are given in (??). A useful expansion of a delta function in terms
of Legendre polynomials is
( 0 ) =

B.6

Pl () Pl (0 ).

(B.80)

l=0

Laguerre Polynomials

Laguerre polynomials are the natural (orthogonal basis) energy weighting functions in which to expand a distribution function in spherical velocity space
see (??). The relevant forms for kinetic theory and plasma physics are defined
in general by
ex
(n + l + 3/2) (x)m
dn 0 x n+l+1/2 1
=
e x
m! (n m)! (m + l + 3/2)
n! xl+1/2 dxn
m=0
(B.81)
in which (x mv 2 /2T = v 2 /vT2 ) is the normalized kinetic energy variable and
l is the integer subscript of the (Legendre) polynomial expansion in spherical
(x) =
L(l+1/2)
n

n
$

DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

12

APPENDIX B. SPECIAL FUNCTIONS

velocity space. These generalized Laguerre polynomials satisfy the differential


equation
(l+1/2)

d2 Ln
dx2

(l+1/2)

+ (l x + 3/2)

dLn
dx

+ n Ln(l+1/2) = 0.

(B.82)

They have a generating function given by


1
exp
(1 z)l+3/2

xz
z1

L(l+1/2)
(x) z n ,
n

n=0

for |z| < 1.

(B.83)

Laguerre polynomials are closely related to Hermite polynomials (their Cartesian velocity space equivalents) and Sonine polynomials (reversed indices, different normalization). The lowest order (n = 0, 1, 2 and l = 0, 1, 2) Laguerre
polynomials are
(1/2)

3
15 5x x2
(1/2)
x, L2

+ ,
=
2
8
2
2

(3/2)

5
35 7x x2
(3/2)
x, L2
=

+ ,
2
8
2
2

(5/2)

7
63 9x x2
(5/2)
x, L2

+ ,
=
2
8
2
2

(1/2)

= 1,

L1

(3/2)

= 1,

L1

(5/2)

= 1,

L1

L0
L0
L0

(B.84)

The orthogonality and values of relevant energy integrals of products of these


Laguerre polynomials are given in (??) and (??).
REFERENCES
Some general compendia of properties of special functions are:
Abromowitz and Stegun, Handbook of Mathematical Functions (1965) [?]
Magnus, Oberhettinger and Soni, Formulas and Theorems for the Special Functions of Mathematical Physics (1966) [?]
Jahnke and Emde, Table of Functions (1945) [?].

DRAFT 22:15
February 2, 2002

c
"J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX C. USEFUL DEFINITE INTEGRALS

Appendix C

Useful Definite Integrals


Definite integrals that often arise in plasma physics are summarized in this
appendix.

C.1

Integrals Involving A Decaying Exponential

Integrals over temporally or spatially decaying processes (e.g., collisional damping at rate 1/ ) often result in integrals of the form
!
!
dt tn et/ = n+1
dx xn ex
(C.1)
0

in which x t/ . The most general definite, dimensionless integral involving


powers of a variable x and the exponential ex is that given by the gamma
(factorial) function, which is defined by Eulers integral:
!
dx xz1 ex , for Re(z) > 0.
(C.2)
(z)
0

Integrating by parts, one obtains the important recursion relation


(z + 1) = z (z).

(C.3)

Using this relation recursively, the gamma function for any argument z > 1 can
be evaluated in terms of (z) for 0 < z 1.
Two values of the argument z of fundamental interest for gamma functions
are z = 1 and z = 1/2. For z = 1 the gamma function becomes simply the
integral of a decaying exponential:
!
dx ex = 1.
(C.4)
(1) =
0

For z = 1/2, using the substitution x = u2 the gamma function becomes the
integral of a Gaussian distribution over an infinite domain:
!

2
du eu = .
(C.5)
( 12 ) = 2
0

DRAFT 10:50
February 3, 2002

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX C. USEFUL DEFINITE INTEGRALS

When the argument of the gamma function is a positive integer (z n > 0),
the gamma function simplifies to a factorial function:
(n + 1) = n (n) = n(n 1) (n 1) = n(n 1)(n 2) 1 n!. (C.6)
Using this factorial form for the gamma function, one thus finds that
!
dt tn et/ = n+1 n!, for n = 0, 1, 2, ,

(C.7)

using the usual convention that 0! 1. The first few of these integrals are


1
1

!
!

1
dt
t/
x
t/
x
1
e
=
dx
e
=
.
(C.8)


0
0
2 2

t /
x
2

When the argument of the gamma function is a positive half integer (z


n + 1/2 > 0), the gamma function simplifies to a double factorial:
(n + 12 )

(n 12 ) (n 12 ) = (n 12 )(n 32 ) (n 32 )

= [(2n 1)(2n 3) 1] ( 12 ) / 2n (2n 1)!!

C.2

/ 2n .
(C.9)

Integrals Over A Maxwellian

When calculating various averages over a Maxwellian distribution, integrals of


the following type occur:
!
!
2
2
2
dv v m ev /vT = vTm+1
du um eu
(C.10)
Im =
0

in which m is a nonnegative integer and in the second, dimensionless integral


u v/vT . This integral can be calculated for arbitrary m 0 by changing
the variable of integration from u to x = u2 = v 2 /vT2 and relating the resulting
integral to the gamma function, (C.2):
Im =

vTm+1
2

dx xm/21/2 ex =

vTm+1
[(m + 1)/2].
2

The integrals for the first few even m [for which (m + 1)/2 becomes
integer and (C.9) applies] are

1
1
1


!
!

2
dv

v 2 /vT
u2
2
2
2
u
v /vT
1/2
=
du
=
e
e

vT
2
0
0
4

4 4

v /vT
u
3/4
DRAFT 10:50
February 3, 2002

c
$J.D
Callen,

(C.11)
a half

(C.12)

Fundamentals of Plasma Physics

APPENDIX C. USEFUL DEFINITE INTEGRALS


The integrals for the first few odd m [for which (m + 1)/2
and (C.6) applies] are

u
v/vT

!
!

2
2
2
dv
u3
v 3 /vT3
du
ev /vT =
eu =

v
T
0
0
5
5 5
v /vT
u

becomes an integer

1/2
1/2
. (C.13)

The natural (orthogonal basis) energy weighting functions for expanding


distribution functions in terms of fluid moments are the Laguerre polynomials
l+1/2
(x), which are defined and discussed in Section B.6. The relevant diLn
mensionless integral of products of Laguerre polynomials that indicates their
orthogonality and normalization is
!
(l+1/2)
dx xl+1/2 ex Ln(l+1/2) (x) Ln!
(x)
0
(C.14)

(l + n + 3/2)

= nn! [2(n + l) + 1]!! n+l+1


= nn!
(n + 1)
2
n!
in which x v 2 /vT2 = mv 2 /2T , and nn! is the Kronecker delta, which is unity
for n = n# and vanishes if n (= n# . The lowest order (n = 0, 1, 2 and l = 0, 1, 2)
integrals of interest are

(1/2)
1/2
1/2
[L0 ]2
[L1 ]2
[L2 ]2
!

(3/2) 2
(3/2)
(3/2)
dx x1/2 ex
x[L1 ]2 x[L2 ]2
x[L0 ]

(5/2) 2

x2 [L0

C.3

1

3/2
2
15/8

(5/2) 2

x2 [L1

(5/2) 2

x2 [L2

3/2

15/4

15/4

105/8 .

105/16

(C.15)

945/32

Integrals Over Sinusoidal Functions

Averaging linear and nonlinear quantities made up of sinusoidally oscillating


components result in integrals of the form
! 2
1
d sinm cosn .
(C.16)
)sinm cosn *
2 0
Trigonometric identities that are useful in reducing these integrals to simpler
forms are
2 sin cos = sin 2,
2 sin2 = (1 cos 2),
2 cos2 = (1 + cos 2),
DRAFT 10:50
February 3, 2002

c
$J.D
Callen,

(C.17)
(C.18)
(C.19)

Fundamentals of Plasma Physics

APPENDIX C. USEFUL DEFINITE INTEGRALS


which are derivable from the more fundamental trigonometric identities
sin (1 + 2 ) = sin 1 cos 2 + cos 1 sin 2 ,
cos (1 + 2 ) = cos 1 cos 2 sin 1 sin 2 .

(C.20)
(C.21)

These last two identities can be also combined to yield


2 sin 1 sin 2 = cos (1 + 2 ) cos (1 2 ),
2 sin 1 cos 2 = sin (1 + 2 ) + sin (1 2 ),
2 cos 1 cos 2 = cos (1 + 2 ) + cos (1 2 ).
Using these trigonometric identities, and the facts that
/ 2
d cos n = 0 for n = 1, 2, , it can be shown that
0

/ 2
0

d sin n = 0 and

0
1/2
.
0
3/8
(C.25)
The natural (i.e., orthogonal basis) functions of sinusoidal functions in which
to expand spherical velocity space latitude angle dependences are the Legendre
polynomials Pl (), which are defined and discussed in Section B.5. The relevant argument of the Legendre polynomials is usually cos . The relevant
integral of products of Legendre polynomials that indicates their orthogonality
and normalization is
!
! 1
2 ll!
(C.26)
d Pl () Pl! () =
d sin Pl (cos ) Pl! (cos ) =
2l
+1
1
0
1
2
sin cos
1
d
sin cos2
2 0
sin2 cos2
!

sin
sin2
sin3
sin4

cos
1
0
cos2
=
0
cos3
4
1/8
cos

(C.22)
(C.23)
(C.24)

0
1/2
0
3/8

in which ll! is the Kronecker delta function which is unity if the indices are
equal and zero otherwise. The first few of these nonvanishing integrals are
2

! 1
! 1
P0
1
2
= 2/3 . (C.27)
cos2
d P12
d(cos )
2
2
2
1
1
P2
(3 cos 1) /4
2/5
REFERENCES
A limited but very useful table of integrals is:
Dwight, Tables of Integrals and Other Mathematical Data (1964) [?]

The most comprehensive tabulation of integrals is provided by:


Gradshteyn and Ryzhik, Table of Integrals, Series and Products (1965) [?]

DRAFT 10:50
February 3, 2002

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX D. VECTOR ANALYSIS

Appendix D

Vector Analysis
The following conventions are used in this appendix and throughout the book:
f, g, , are scalar functions of x, t;
A, B, C, D
are vector functions of x, t;
A = |A| A A is the magnitude or length of the vector A;

eA A/A is a unit vector in the A direction;


x is the vector from the origin to the point (x, y, z);
T, W, AB, etc., are dyad (second rank tensor) functions of x, t that will
be called simply tensors;
I is the identity tensor or unit dyad;
TT is the transpose of tensor T (interchange of indices of the tensor
elements), a tensor;
tr(T) is the trace of the tensor T (sum of its diagonal elements), a scalar;
det(T) #T# is the determinant of the tensor T (determinant of the
matrix of tensor elements), a scalar.

D.1

Vector Algebra

Basic algebraic relations:


A + B = B + A, commutative addition
A + (B + C) = (A + B) + C, associative addition

(D.1)
(D.2)

A B = A + (B), difference
f A = Af, commutative scalar multiplication

(D.3)
(D.4)

(f + g)A = f A + gA, distributive scalar multiplication


f (A + B) = f A + f B, distributive scalar multiplication

(D.5)
(D.6)

f (gA) = (f g)A,

(D.7)

DRAFT 11:26
October 11, 2002

associative scalar multiplication

c
%J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX D. VECTOR ANALYSIS


Dot product:
A B = 0 implies A = 0 or B = 0, or A B
A B = B A, commutative dot product

A (B + C) = A B + A C, distributive dot product


(f A) (gB) = f g(A B), associative scalar, dot product

(D.8)
(D.9)
(D.10)
(D.11)

Cross product:
AB = 0 implies A = 0 or B = 0, or A # B
AB = BA, AA = 0,
A(B + C) = AB + AC,
(f A)(gB) = f g(AB),

anti-commutative cross product


distributive cross product

associative scalar, cross product

(D.12)
(D.13)
(D.14)
(D.15)

Scalar relations:
A BC = AB C = (CA) B,

dot-cross product

(AB) (CD) = (A C)(B D) (A D)(B C)


(AB) (CD) + (BC) (AD) + (CA) (BD) = 0

(D.16)
(D.17)
(D.18)

Vector relations:
A(BC) = B(A C) C(A B),

bac cab rule

= (CB)A = A (CB BC)


A(BC) + B(CA) + C(AB) = 0

(D.19)
(D.20)

(A B)C = A (BC), associative dot product


(AB)(CD) = C(AB D) D(AB C)

(D.21)

= B(CD A) A(CD B)

(D.22)

Projection of a vector A in directions relative to a vector B:


+ A
A = A! (B/B) + A = A! b
B/B, unit vector in B direction
b
A, component of A along B
A! B A/B = b
A B(BA)/B 2 ,
bA)

= b(

D.2

(D.23)
(D.24)
(D.25)

component of A perpendicular to B
(D.26)

Tensor Algebra

Scalar relations:
I : AB (I A) B = A B
AB : CD A (B C)D = (B C)(A D)
DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

(D.27)

Fundamentals of Plasma Physics

APPENDIX D. VECTOR ANALYSIS

dot!product

cross!product

dot-cross!product

C
B
A

Figure D.1: Schematic illustration of dot, cross and dot-cross products of vectors.

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX D. VECTOR ANALYSIS


= D AB C = B CD A

(D.28)
(D.29)

I : T = tr(T), T : T |T|
T : AB = (T A) B = B T A
AB : T = A (B T) = B T A
BT : W = (T W)T : BI
2

(D.30)
(D.31)
(D.32)

Vector relations:
IA = AI = A
A TT = T A, TT A = A T
A (CB BC) = A(BC)
= B(A C) C(A B),
(AC) T = A (CT) = C (AT)

(D.33)
(D.34)
bac cab rule

T (AC) = (TA) C = (TC) A


A (TC) = (A T)C = C(A T)

(D.35)
(D.36)
(D.37)
(D.38)

(AT) C = A(T C) = (T C)A


A (TC) C (TA) = [I tr(T) T] (AC)

(AT) C (CT) A = (AC) [I tr(T) T]

(D.39)
(D.40)
(D.41)

Tensor relations:
I AB = (I A)B = AB,
IA = IA

AB I = A(B I) = AB

A(BC) = (AB)C, (AB)C = A(BC)


(AB)I = I(AB) = BA AB
(AT)T = TT A, (TA)T = ATT
(AT) (AT)T = I[A tr(T) T A]
(TA) (TA)T = I[A tr(T) A T]
TS = 12 (T + TT ), symmetric part of tensor T
TA = 12 (T TT ), anti-symmetric part of tensor T
BTS B = B 2 TS (BB TS + TS BB)
(IB 2 BB)(IB 2 BB) TS /B 2 BB(BB TS )/B 2

D.3

(D.42)
(D.43)
(D.44)
(D.45)
(D.46)
(D.47)
(D.48)
(D.49)
(D.50)
(D.51)

Derivatives

Temporal derivatives:
dA
is a vector tangent to the curve defined byA(t)
dt
df
dA
d
(f A) = A + f
dt
dt
dt
DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

(D.52)
(D.53)

Fundamentals of Plasma Physics

APPENDIX D. VECTOR ANALYSIS


dA dB
d
(A + B) =
+
dt
dt
dt
dA
dB
d
(A B) =
B+A
dt
dt
dt
dA
dB
d
(AB) =
B + A
dt
dt
dt

(D.54)
(D.55)
(D.56)

Definitions of partial derivatives in space ( /x = del or nabla is the


differential vector operator):
f
, gradient of scalar function f , a vector vector in direction
x
of and measure of the greatest rate of spatial change of f
(D.57)

A, divergence of vector function A, a scalar


x
divergence ( A > 0) or convergence ( A < 0) of A lines

(D.58)

A, curl (or rotation) of vector function A, a vector1


x
vorticity of A lines
(D.59)

2 f f, del square or Laplacian (divergence of gradient)


derivative of scalar function f , a scalar, which is sometimes
written as f three-dimensional measure of curvature of f
(f is larger where 2 f < 0 and smaller where 2 f > 0)
2 A ( )A = ( A) (A),
of vector function A, a vector

Laplacian derivative
(D.61)

ey + z
ez and |x|
For the general vector coordinate x x
ex + y
x = 3, (x/|x|) = 2/|x|
x = 0, (x/|x|) = 0
|x| = x/|x|,
x = I

(D.60)

(1/|x|) = x/|x|3

(A )(x/|x|) = [A (x A)x/|x|2 ]/|x| A /|x|


2 (1/|x|) (1/|x|) = (x/|x|3 ) = 4(x)

!
x2 + y 2 + z 2 :

(D.62)
(D.63)
(D.64)
(D.65)
(D.66)
(D.67)

1 Rigorously speaking, the cross product of two vectors and the curl of a vector are pseudovectors because they are anti-symmetric contractions of second rank tensors see tensor
references at end of this appendix.

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX D. VECTOR ANALYSIS


First derivatives with scalar functions:
(f + g) = f + g

(D.68)

f A = f A + f A
f A = f A + f A

(D.71)
(D.72)

f T = f T + f T

(D.74)

(f g) = (f )g + f g = (gf )
(f A) = (f )A + f A

(D.69)
(D.70)

f T = f T + f T

(D.73)

First derivative scalar relations:


(A + B) = A + B

(D.75)

CB : A + AB : C
A B C C B A (CA AC) : B = (AC) B

(D.77)
(D.78)

+ (BA) (C) + (AC) (B)


I : B = B

(D.79)
(D.80)

(AB) = B A A B
(B )(A C) = C (B )A + A (B )C

2A B C 2CA : B = A (B C) + C (B A)
B (A C) + (BC) (A)

AI : B = A B
A T = (A T) A : T = (A T) T : A

(D.76)

(D.81)
(D.82)

First derivative vector relations:


(A + B) = A + B
(A B) = A(B) + B(A) + (A )B + (B )A

= (A) B + (B) A
(B /2) (B B/2) = B(B) + (B )B = (B) B
(B )(AC) = (B )AC + A(B )C
2

(D.83)
(D.84)
(D.85)
(D.86)

AB = ( A)B + (A )B = ( A)B + A (B)

(D.87)

A(B) = (B) A A (B) = (B) A (A )B


(AB) = A( B) B( A) + (B )A (A )B

(D.90)

A BC C BA = [( B)I B] (AC)
AB C CB A = (AC) [( B)I B]

(D.93)
(D.94)

I = 0
(IA) = A

(D.88)
(D.89)

= (BA AB)
A BC + CB A = C[A(B)]

(D.91)
(D.92)

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX D. VECTOR ANALYSIS


First derivative tensor relations:
I B = B, B I = B
AB = (A)B AB

(D.95)
(D.96)

(AB) = AB BA
AB + BA

(D.97)

= I[( B)A (B) A] + [A (B)]I A(B)


= I[( B)A A (B)] + [A (B)]I (B)A
BA + (AB)T = [A (B)]I A(B)
AB + (BA)T = [A (B)]I (B)A
AB (AB)T = I[( B)A (B) A]
BA (BA)T = [( B)A A (B)]I

(D.98)
(D.99)
(D.100)
(D.101)
(D.102)

Second derivative relations:


f 2 f
f = 0
f g = 0
A 2 A = ( A) (A)
A = 0

(B )A = (B )( A) (A) (B)
[(A )A]

= (A )(A) + ( A)(A) [(A) ]A

(D.103)
(D.104)
(D.105)
(D.106)
(D.107)
(D.108)
(D.109)

+ A ,
B/B, A = A! b
Derivatives of projections of A in B direction [b
A, A b(
bA),

)b
= b(

]:
A! b
(b
b)
A = (A! /B)( B) + (B )(A! /B) + A
)b
] (1/B) b
(BA)
A = A [ ln B + (b
)b

A b
b
b
: A = (b
)A! A (b
b

(D.110)
(D.111)

(BA) + A (A! /B)( B) (D.112)


= A ln B (1/B)b
(BA ) = (b
f )(b
b)
(D.113)
For A = (1/B 2 ) Bf, b

D.4

Integrals

For a volume V enclosed by a closed, continuous surface S with differential


dS where n
is the
volume element d3x and differential surface element dS n
unit normal outward from the volume V , for well-behaved functions f, g, A, B
and T:
""
"
d3x f = % dS f,
(D.114)
V

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX D. VECTOR ANALYSIS


""
d x A = % dS A,
divergence or Gauss theorem,
"" S
"V
d3x T = % dS T,
""S
"V
3
d x A = % dSA,
" S
""
"V
3
2
3
d x f g =
d x f g + % dS f g,
"

"
"

(D.115)
(D.116)
(D.117)

Greens first identity,


""
d3x (f 2 g g2 f ) = % dS (f g gf ),

(D.118)

Greens second identity,

d3x [A (B) B (A)]


""
= % dS [B(A) A(B)],

(D.119)

vector form of Greens second identity.

(D.120)

The gradient, divergence and curl partial differential operators can be defined
using integral relations in the limit of small surfaces S encompassing small
volumes V , as follows:
$
#
""
1
% dS f
gradient,
(D.121)
f lim
V 0 V
S
$
#
""
1
% dS A
divergence,
(D.122)
A lim
V 0 V
S
#
$
""
1
% dSA
A lim
curl.
(D.123)
V 0 V
S

For S representing an open surface bounded by a closed, continuous contour C


with line element d" which is defined to be positive when the right-hand-rule
sense of the line integral around C points in the dS direction:
%
""
dSf =
d"f,
(D.124)
C
%
""S
dS A =
d" A,
Stokes theorem, (D.125)
S
C
%
""
(dS)A =
d"A,
(D.126)
C
%
%
%
""S
dS (f g) =
d" f g =
f dg =
g df,
S

Greens theorem. (D.127)

The appropriate differential line element d", surface area dS, and volume d3x
can be defined in terms of any three differential line elements d"(i), i = 1, 2, 3
DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX D. VECTOR ANALYSIS


that are linearly independent [i.e., d"(1) d"(2)d"(3) (= 0] by
d" = d"(i), i = 1, 2, or 3,
dS = d"(i)d"(j),
d3x = d"(1) d"(2)d"(3),

differential line element,


differential surface area,
differential volume.

(D.128)
(D.129)
(D.130)

In exploring properties of fluids and plasmas we often want to know how


the differential line, surface and volume elements change as they move with the
fluid flow velocity V. In particular, when taking time derivatives of integrals,
we need to know what the time derivatives of these differentials are as they
are carried along with a fluid. To determine this, note first that if the flow is
uniform then all points in the fluid would be carried along in the same direction
at the same rate; hence, the time derivatives of the differentials would vanish.
However, if the flow is nonuniform, the differential line elements and hence all
the differentials would change in time. To calculate the time derivatives of the
differentials, consider the motion of two initially close points x1 , x2 as they are
carried along with a fluid flow velocity V(x, t). Using the Taylor series expansion
V(x2 , t) = V(x1 , t)+(x2 x1 ) V+ and integrating the governing equation
dx/dt = V over time, we obtain
" t
dt$ (x$2 x$1 ) V +
(D.131)
x2 x1 = x2 (t = 0) x1 (t = 0) +
0

in which x2 (t = 0) and x1 (t = 0) are the initial positions at t = 0. Taking the


time derivative of this equation and identifying the differential line element d"
as x2 x1 in the limit where the points x2 and x1 become infinetesimally close,
we find
d
(D.132)
d" (d") = d" V.
dt
The time derivative of the differential surface area dS can be calculated by
taking the time derivative of (D.129) and using this last equation to obtain
d

(dS) = d"(1)d"(2)
+ d"(1)d"(2)
dt
= d"(1) Vd"(2) d"(2) Vd"(1)

dS

( V) dS V dS

(D.133)

in which (D.93) and (D.33) have been used in obtaining the last form. Similarly,
the time derivative of the differential volume element moving with the fluid is
d 3
(d x)
dt

= d"(3)
dS + d"(3) dS

= d"(3) V dS + d"(3) ( V)dS d"(3) V dS


(D.134)
= ( V) d3x,

which shows that the differential volume in a compresssible fluid increases or


decreases according to whether the fluid is rarefying ( V > 0) or compressing
( V < 0).
DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

10

APPENDIX D. VECTOR ANALYSIS

D.5

Vector Field Representations

Any vector field B can be expressed in terms of a scalar potential M and a


vector potential A:
B = M + A,

potential representation.

(D.135)

The M part of B represents the longitudinal or irrotational (M = 0)


component while the A part represents the transverse or solenoidal component ( A = 0). A vector field B that satisfies B = 0 is called a longitudinal or irrotational field; one that satisfies B = 0 is called a solenoidal
or transverse field. For a B(x) that vanishes at infinity, the potentials M and
A are given by Greens function solutions
"
"
( B)x!
(B)x!
,
A(x)
=
d3x$
.
(D.136)
M (x) = d3x$
4|x x$ |
4|x x$ |
When there is symmetry in a coordinate (i.e., a two or less dimensional
system), a solenoidal vector field B can be written in terms of a stream function
in such a way that it automatically satisfies the solenoidal condition B = 0:
B = = ||
e = ,
For this situation the vector potential becomes

stream function form.


(D.137)

A = = ||
e .

(D.138)

For a fully three-dimensional situation with no symmetry, a solenoidal vector


field B can in general be written as
B = ,

Clebsch representation,

(D.139)

In this representation and are stream functions that are constant along the
vector field B since B = 0 and B = 0.

D.6

Properties Of Curve Along A Vector Field

The motion of a point x along a vector field B is described by


B
dx

=
=b
T,
d)
B

tangent vector

(D.140)

is tangent to
in which d) is a differential distance along B. The unit vector b
T a unit tangent
the vector field B(x) at the point x and so is often written as
vector.
The curvature vector of the vector field B is defined by

db
d2 x
)b
= b(

= (b
b),
=
2
d)
d)

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

curvature vector

(D.141)

Fundamentals of Plasma Physics

11

APPENDIX D. VECTOR ANALYSIS

in which (D.85) has been used in the obtaining the last expression. The unit
vector in the curvature vector direction is defined by
)b
/ |(b
)b|,

(b

curvature unit vector.

(D.142)

The local radius of curvature vector RC is in the opposite direction from the
curvature vector and is defined by
RC /||2 ,

2
= RC /RC
,

radius of curvature.

(D.143)

Hence, |RC | RC = 1/|| is the magnitude of the local radius of curvature


the radius of the circle tangent to the vector field B(x) at the point x.
A triad of orthogonal unit vectors (see Fig. D.2) can be constructed from the
T and an arbitrary unit vector
N normal (or perpendicular)
tangent unit vector
to the vector field B(x) at the point x:

N,

T b,
N and
B
T
N = b

Frenet unit vector triad

(D.144)

B is the binormal unit vector, the third orthogonal unit vector. The
in which
component of a vector C in the direction of the vector field B is called the
C. The component in the
TC = b
N direction is
parallel component: C!
N C. The component in the
B direction,
called the normal component: CN
which is perpendicular to the
T
N plane, is called the binormal component:
BC =
T
N C.
CB
Consider for example the components of the curvature vector . Since
= 0, the curvature vector has no parallel component (! = 0) the
b
curvature vector for the vector field B(x) is perpendicular to it at the point
x. The components of the curvature vector relative to a surface (x) = constant in which the vector field lies (i.e., B = 0) can be specified as follows.
N /||.
Define the normal to be in the direction of the gradient of :
Then, the components of the curvature vector perpendicular to (normal) and
lying within (geodesic) the surface are given by
N = /||,
n =

B = (b)
/|b|,
g =

normal curvature,

(D.145)

geodesic curvature.

(D.146)

The torsion (twisting) of a vector field B is defined by


B
d
)(b
N),
= (b
d)

torsion vector.

(D.147)

The binormal component of the torsion vector vanishes (B


B = 0). The
normal component of the torsion vector locally defines the scale length L along
the vector B over which the vector field B(x) twists through an angle of one
radian:
L 1/|N |,

N
N

DRAFT 11:26
October 11, 2002

d
B
)(b
N),
=
N (b
d)
c
%J.D
Callen,

torsion length. (D.148)

Fundamentals of Plasma Physics

12

APPENDIX D. VECTOR ANALYSIS

curvature

torsion

B
B

Rc

shear

Figure D.2: Properties (curvature, torsion, shear) of a spatially inhomogeneous


B/B = dx/d) =
vector field B(x). The unit vector b
T is locally tangent to
N is perpendicular to the vector field B,
the vector field B. The unit normal

B is orthogonal to both b
shown here in the curvature direction. The binormal
N.
and

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

13

APPENDIX D. VECTOR ANALYSIS

N is taken to be in the direction, the parallel comIf the unit normal


=
ponent of the torsion vector is equal to the geodesic curvature [! b

)b
/ |b|

(b)
(b
g ].
The local shear (differential twisting motion, or nonplanar differential tan=
T and
N) in a vector field B is given
gential motion in the plane defined by b
by the binormal component of the curl or rotation in the binormal unit vector:

N) (b
N) 1/LS ,
B
B = (b

local shear.

(D.149)

The shear length LS is defined as the scale length over which the vector field
B(x) shears through an angle of one radian. The parallel component of the
total curl or rotation of a vector field B is given by a combination of its torsion
N
N:
and shear, and
N) 2
)(b
N) +
b
= (b
N) (b
N (b
N
N
b

= + 2N + N
N,
total rotation in B field.
(D.150)
N is taken to be in the direction,
N
N = 0 and then
If the normal

1
(b)
(b)

,
2

LS
|b|

and

local shear with


N /||,
(D.151)

b
= + 2N .
b

(D.152)

In the absence of shear ( = 0), this last relation yields N = (1/2)b


the torsion for rigid body rotation is just half the parallel component of the
rotation in the vector field B.
N is taken to be in the curvaIn most applied mathematics books the normal
N
) instead of the direction. Then, the parallel
ture vector direction (i.e.,
= b
(b
)b
=
component of the torsion vector also vanishes [! b

b
= 0] and
N,
for
N
.
(D.153)
N

For this case the interrelationships between the triad of unit vectors
T,
N,
B are
given by the Frenet-Serret formulas:
T
d

= N
N,
T B/B b,
d)
N
d
)b
/ |(b
)b|,

= N
T + N
B,
N
= (b
d)
d
B
.

= N
N,
B
T
N = b
d)

(D.154)

The local shear and total rotation in the vector field B for this case are as
N.
given above in (D.149) and (D.150), respectively, for a general unit normal

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

14

APPENDIX D. VECTOR ANALYSIS

D.7

Base Vectors and Vector Components

The three vectors e1 , e2 , e3 , which are not necessarily orthogonal, can be used as
a basis for a three-dimensional coordinate system if they are linearly independent
(i.e., e1 e2 e2 (= 0). The three reciprocal base vectors e1 , e2 , e3 are defined by
where
ji

&

ei ej = ji ,
1,
0,

i = j,
i (= j,

(D.155)

Kronecker delta.

(D.156)

The reciprocal base vectors can be written in terms of the original base vectors:
e1 =

e1

e2 e3
e3 e1
e1 e2
, e2 = 1
, e3 = 1
.
2
3
2
3
(e e )
e (e e )
e (e2 e3 )

(D.157)

Or, in general index notation


ei = .ijk

ej ek
,
e1 (e2 e3 )

i, j, k = permutations of 1, 2, 3

(D.158)

in which

+1 when i, j, k is an even permutation of 1, 2, 3


1 when i, j, k is an odd permutation of 1, 2, 3
.ijk =

0 when any two indices are equal


Levi-Civita symbol.
(D.159)
The reciprocal LeviCivita symbol .ijk is the same, i.e., .ijk =.ijk . These formulas are also valid if the subscripts and subscripts are reversed. Thus, the
original base vectors could be the reciprocal base vectors ei and the reciprocal base vectors could be the original base vectors ei since both sets of base
vectors are linearly independent. Either set can be used as a basis for representing three-dimensional vectors.
The identity tensor can be written in terms of the base or reciprocal vectors
as follows:
* i
1
2
3
I
i e ei = e e1 + e e2 + e e3
identity tensor
(D.160)
*
i
1
2
3

i ei e = e1 e + e2 e + e3 e .
This definition can be used to write any vector or operator in terms of either
its base or reciprocal vector components:
+
Ai ei , Ai A ei ,
A = A I = (A e1 )e1 + (A e2 )e2 + (A e3 )e3 =
+
i
Aj ej , Aj A ej ,
= (A e1 )e1 + (A e2 )e2 + (A e3 )e3 =
j

I = e1 (e1 ) + e2 (e2 ) + e3 (e3 )


= e1 (e1 ) + e2 (e2 ) + e3 (e3 ).
DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

(D.161)

(D.162)

Fundamentals of Plasma Physics

15

APPENDIX D. VECTOR ANALYSIS

The dot product between two vectors A and B is given in terms of their
base and reciprocal vector components by
+
+
+
+
Ai Bi =
Ai B i =
(ei ej )Ai B j =
(ei ej )Ai Bj . (D.163)
AB =
i

ij

ij

Similarly, the cross product between two vectors is given by


+
+
Ai B j ei ej =
Ai B j ek (e1 e2 e3 )
AB =
=
, 1
, e
, 1
= (e1 e2 e3 ) ,
, A1
, B

ij

ijk

ij

Ai Bj ei ej =

e2
A2
B2

ijk

.ijk Ai Bj ek (e1 e2 e3 )

,
,
, e1
e3 ,
,
,
1
2
3 ,
A3 ,
= (e e e ) , A1
,
, B1
B3 ,

e2
A2
B2

e3
A3
B3

,
,
,
, . (D.164)
,
,

The dot-cross product of three vectors is given by


+
+
Ai B j C k ei ej ek =
.ijk Ai B j C k (e1 e2 e3 )
A BC =
=

ijk

ijk

ijk

, 1
, A
, 1
= (e1 e2 e3 ) ,
, B1
, C

Ai Bj Ck ei ej ek =
2

A
B2
C2

ijk

.ijk Ai Bj Ck (e1 e2 e3 )

,
,
, A1
A ,
,
,
1
2
3 ,
B3 ,

e
e
)
=
(e
, B1
,
, C1
C3 ,
3

A2
B2
C2

A3
B3
C3

,
,
,
, . (D.165)
,
,

For the simplest situation where the three base vectors e1 , e2 , e3 are orthogonal (e1 e2 = e2 e3 = e1 e3 = 0), the reciprocal vectors point in the same
directions as the original base vectors. Thus, after normalizing the base and
reciprocal vectors they become equal:
1 = e1 /|e1 |
1 = e1 /|e1 | = e
e
2 = e2 /|e2 |
2 = e2 /|e2 | = e
e
3 = e3 /|e3 |
3 = e3 /|e3 | = e
e

orthogonal
unit
vectors.

(D.166)

The simplifications of (??)(??) are given in (D.196)(D.201) in the section


(D.9) below on orthogonal coordinate systems.

D.8

Curvilinear Coordinate Systems

Consider transformation from the Cartesian coordinate system x = (x, y, z)


to a curvilinear coordinate system labeled by the three independent functions
u1 , u2 , u3 :
x = x(u1 , u2 , u3 ) :

DRAFT 11:26
October 11, 2002

x = x(u1 , u2 , u3 ),

y = y(u1 , u2 , u3 ),

c
%J.D
Callen,

z = z(u1 , u2 , u3 ).
(D.167)

Fundamentals of Plasma Physics

16

APPENDIX D. VECTOR ANALYSIS

The transformation is invertible if the partial derivatives x/ui for i = 1, 2, 3


are continuous and the Jacobian determinant (i.e., x/u1 x/u2 x/u3 )
formed from these nine partial derivatives does not vanish in the domain of
interest. The inverse transformation is then given by
ui = ui (x) :

u1 = u1 (x, y, z),

u2 = u2 (x, y, z),

u3 = u3 (x, y, z). (D.168)

In a curvilinear coordinate system there are three coordinate surfaces:


u1 (x) = c1
u2 (x) = c2
u3 (x) = c3

(u2 , u3 variable),
(u1 , u3 variable),
(u1 , u2 variable).

(D.169)

There are also three coordinate curves given by


u2 (x) = c2 ,
u3 (x) = c3 ,
u1 (x) = c1 ,

u3 (x) = c3
u1 (x) = c1
u2 (x) = c2

(u1 variable),
(u2 variable),
(u3 variable).

(D.170)

The direction in which ui increases along a coordinate curve is taken to be the


positive direction for ui . If the curvilinear coordinate curves intersect at right
angles (i.e., ui uj = 0 except for i = j), then the system is orthogonal. The
familiar Cartesian, cylindrical and spherical coordinate systems are all orthogonal. They are discussed at the end of the next section which covers orthogonal
coordinates.
A nonorthogonal curvilinear coordinate system can be constructed from an
invertible set of functions u1 (x), u2 (x), u3 (x) as follows. A set of base vectors
ei can be defined by
ei = ui ,

i = 1, 2, 3 contravariant base vectors.

(D.171)

These so-called contravariant (superscript index) base vectors point in the direction of the gradient of the curvilinear coordinates ui , and hence in the directions
perpendicular to the ui (x) = ci surfaces. The set of reciprocal base vectors ei
is given by
ei = .ijk

ej ek
.ijk
= 1 uj uk ,
e2 e3
J

e1

covariant base vectors,

(D.172)

in which
J 1 u1 u2 u3 = e1 e2 e3

inverse Jacobian

(D.173)

is the Jacobian of the inverse transformation from the ui curvilinear coordinate system back to the original Cartesian coordinate system.
An alternative form for the reciprocal base vectors can be obtained from the
definition of the derivative of one of the curvilinear
coordinates ui (x) in terms
*
i
i
i
of the gradient: du = u dx = u j (x/uj ) dxj , which becomes an
DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

17

APPENDIX D. VECTOR ANALYSIS

identity if and only if ui (x/uj ) = ji . Since this last relation is the same
as the defining relation for reciprocal base vectors (ei ej = ji ), it follows that
x
, i = 1, 2, 3 covariant base vectors.
(D.174)
ui
The so-called covariant (subscript index) base vectors point in the direction
of the local tangent to the ui variable coordinate curve (from the x/ui definition), i.e., parallel to the ui coordinate curve. Alternatively, the covariant
base vectors can be thought of as pointing in the direction of the cross product
of contravariant base vectors for the two coordinate surfaces other than the ui
coordinate being considered (from the uj uk definition). That these two
directional definitions coincide follows from the properties of curvilinear surfaces and curves. The contravariant base vectors ei can also be defined as the
reciprocal base vectors of covariant base vectors ei :
ei =

ei = .ijk

ej ek
.ijk x
x
=

;
e1 e2 e2
J uj uk

i, j, k = permutations of 1, 2, 3
contravariant base vectors
(D.175)

in which

x
x x

= e1 e2 e3 Jacobian
(D.176)
u1 u2 u3
is the Jacobian of the transformation from the Cartesian coordinate system to
the curvilinear coordinate system specified by the functions ui .
The geometrical properties of a nonorthogonal curvilinear coordinate system
are characterized by the dot products of the base vectors:
J=

gij ei ej =
g

ij

x x

ui uj

e e = u u
i

covariant metric elements,


j

(D.177)

contravariant metric elements.

These symmetric tensor metric elements can be used to write the covariant
components of a vector in terms of its contravariant components and vice versa:
+
+
(A ej )(ej ei ) =
gij Aj
Ai A ei = A I ei =
j

A Ae = AIe =
i

+
j

(A ej )(e e ) =
j

g ij Ai .

(D.178)

Similarly, they can also be used to write the covariant base vectors in terms of
the contravariant base vectors and vice versa:
+
+
gij ej , ei =
g ij ei .
(D.179)
ei =
j

From the dot product between these relations and their respective reciprocal
base vectors it can be shown that
+
+
gij g jk =
g kj gji = ik .
(D.180)
j

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

18

APPENDIX D. VECTOR ANALYSIS

The determinant of the matrix comprised of the metric coefficients is called


g:

, ,1
g #gij # = ,g ij , ,

(D.181)

in which the second relation follows from interpreting the summation relations
at the end of the preceding paragraph in terms of matrix operations: [gij ][g ik ]
= [I], which yields [gij ] = [g jk ]1 . Since the determinant of the inner product
of two matrices is given by the product of the determinants of the two matrices,
,
, ,
,,
, #
$2
, x x , , x , , x ,
,=,
,,
, = x x x

= J 2.
g = #gij # = ,
, ui uj , , ui , , uj ,
u1 u2 u3
(D.182)
Thus, the determinant of the metric coefficients is related to the Jacobian and
inverse Jacobian as follows:
J=
J

g = e1 e2 e3 =

x x
x

u1 u2 u3

Jacobian,

= 1/ g = e e e = u u u
1

(D.183)

inverse Jacobian.

The various partial derivatives in space can be worked out in terms of covariant derivatives (/ui ) using the properties of the covariant and contravariant
base vectors for a general, nonorthogonal curvilinear coordinate system as follows:
f =

+
i

ui

+ f
f
=
ei i
i
u
u
i

gradient,

(D.184)
+
ei
ei

g (A ei ) =
A = (A I) =
( gAi )
g
g
i
i
+1
+ 1
( g A ei ) =
(J A ui )
divergence,
=

i
i
g
u
J
u
i
i
(D.185)
+
+
(A ej )ej =
Aj uj
A = (A I) =
+

+ Aj

+ .ijk (A ej )
=
ui uj =

i
u
g
ui
ij
ijk
,
,
, e1
e2
e3 ,
,
,
1

,
= ,
1
2
3 ,
,
u
u
u
g,
A1 A2 A3 ,

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

curl,
(D.186)

Fundamentals of Plasma Physics

19

APPENDIX D. VECTOR ANALYSIS


f =

2 f

+ 1
+ f
( g ei
ej j )

i
g
u
u
i
j

+1
+ 1
f
f
( g g ij j ) =
(J ui uj j )

i
i
g u
u
J u
u
ij
ij

Laplacian.

(D.187)
Differential line, surface and volume elements can be written in terms of
differentials of the coordinates ui of a general, nonorthogonal curvilinear coordinate system as follows. Total vector differential and line elements are:
+
+ x
dxi =
ei dxi
i
u
i
i
-*

i
j
|d"| dx dx =
ij gij du du

dx =

(D.188)
metric of coordinates.

Differential line elements d"(i) along curve ui (duj = duk = 0) for i, j, k =


permutations of 1, 2, 3 are
.ijk
d"(i) = ei dui = uj uk dui
g

|d"(i)| = ei ei dui = gii dui

(D.189)

The differential surface element dS(i) in the ui = ci surface (dui = 0) for i, j, k


= permutations of 1, 2, 3 is

dS(i) d"(j)d"(k) = g .ijk ui duj duk


!
2 duj duk =
|dS(i)| = gjj gkk gjk
g ii g duj duk

(D.190)

The differential volume element is

d3x d"(1) d"(2)d"(3) = e1 (e2 e3 ) du1 du2 du3 =

D.9

g du1 du2 du3 .


(D.191)

Orthogonal Coordinate Systems

Consider transformation from the Cartesian coordinate system x = (x, y, z)


to an orthogonal curvilinear coordinate system defined by three independent
functions ui = ui (x, y, z) for i = 1, 2, 3. [Here, the superscripts 1,2,3 are not
powers; rather, they represent labels for the three functions. The functions are
labeled in this way to maintain consistency with the general (nonorthogonal)
curvilinear coordinate literature.] The coordinate surfaces are defined by ui =
ci , where ci are constants. The three orthogonal unit vectors that point in
directions locally perpendicular to the coordinate surfaces are

ei ui /|ui | orthogonal unit vectors.


DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

(D.192)

Fundamentals of Plasma Physics

20

APPENDIX D. VECTOR ANALYSIS

For the simplest orthogonal coordinate system, the Cartesian coordinate system,
, e
2 = y = y
, e
3 = z = z
.
1 = x = x
e
Because of the normalization and assumed orthogonality of these unit vectors,
&
1, for i = j,
ej = ij
Kronecker delta.
(D.193)

ei
0, for i (= j,

The cross products of unit vectors are governed by the right-hand rule which is
embodied in the mathematical relation
ej = .ijk
ek

ei

(D.194)

in which the Levi-Civita symbol .ijk is defined by

+1, for i, j, k = 1, 2, 3 or 2, 3, 1 or 3, 1, 2 (even permutations)


1, for i, j, k = 2, 1, 3 or 1, 3, 2 or 3, 2, 1 (odd permutations)
.ijk

0, for any two indices the same.


(D.195)
A vector A can be represented in terms of its components in the orthogonal
ei :
directions (parallel to ui ) of the unit vectors
+
Ai
ei = A1
e1 + A2
e2 + A3
e3 , Ai A
ei
(D.196)
A=
i

For an orthogonal coordinate system the identity dyad or tensor is


+

ei
ei =
e1
e1 +
e2
e2 +
e3
e3 identity tensor.
I=

(D.197)

Thus, the vector differential operator becomes


+
1 (
2 (
3 (

ei (
ei ) = e
e1 ) + e
e2 ) + e
e3 )
= I =
i

+
i

ui

= u1 1 + u2 2 + u3 3 .
ui
u
u
u

(D.198)

Here and below the sum over i is over the three components 1,2,3.
Using the relations for the dot and cross products of the unit vectors
ei given
in (D.193) and (D.194) the dot, cross and dot-cross products of vectors become
+
Ai Bi = A1 B1 + A2 B2 + A3 B3 ,
(D.199)
AB =
i
,
,
,
e2
e3 ,
+
+
, e1
,
,
A
A
A
AB =
Ai Bj
ei
ej =
.ijk Ai Bj
ek = ,
2
3 ,
, 1
, B1 B2 B3 ,
ij
ijk
=
e1 (A2 B3 A3 B2 ) +
e2 (A3 B1 A1 B2 ) +
e3 (A1 B2 A2 B1 ). (D.200)
,
,
, A1 A2 A3 ,
+
,
,
,
.ijk Ai Bj Ck = ,
A BC =
(D.201)
, B1 B2 B3 , .
,
C1 C2 C3 ,
ijk

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

21

APPENDIX D. VECTOR ANALYSIS


The differential line element in the ith direction is given by
d"(i) =
ei hi dui , with hi 1/|ui |,

differential line element.

(D.202)

Thus, the differential surface vector for the ui = ci surface, which is defined by
dS(i) = d"(j)d"(k), becomes
dS(i) =
ei hj hk duj duk , for i (= j (= k,

differential surface area.

(D.203)

Since the differential volume element is d3x = d"(i) dS(i) = d"(1) d"(2)d"(3)
and the Jacobian of the transformation is given by J = 1/(u1 u2 u3 )
= h 1 h2 h 3 ,
(D.204)
d3x = h1 h2 h3 du1 du2 du3 , differential volume.
are

For orthogonal coordinate systems the various partial derivatives in space


+
+
ei f
=

ei (
ei ) f,
(D.205)
i
hi u
i
i
$ +
#
$
+ 1 #J
1

h 1 h2 h 3
A
ei =
A
ei , (D.206)
A =
J ui hi
h1 h2 h3 ui
hi
i
i
f =

A =

+ .ijk hk
+ .ijk hk
ek
ek
(hj A
ej ) =
(hj A
ej ), (D.207)
i
J
u
h1 h2 h3 ui
ijk

+1
f=
J ui
i
2

ijk

J f
h2i ui

+
i

h1 h2 h3 ui

h1 h2 h3 f
h2i ui

(D.208)

The three most common orthogonal coordinate systems are the Cartesian,
cylindrical, and spherical coordinate systems. Their coordinate surfaces and
unit vectors are shown in Fig. D.3. They will be defined in this book by
Cartesian : ui = (x, y, z)
hx = 1, hy = 1, hz = 1

J = 1;

cylindrical : ui = (r, , z)
!
r x2 + y 2 , arctan(y/x),

z z,

x = r cos , y = r sin , z = z,
hr = 1, h = r, hz = 1 =

J = r;

(D.209)

(D.210)

spherical : ui = (r, , )
!
!
r x2 + y 2 + z 2 , arctan( x2 + y 2 /r), arctan(y/x),
x = r sin cos , y = r sin sin , z = r cos ,
(D.211)
hr = 1, h = r, h = r sin = J = r2 sin .
DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

22

APPENDIX D. VECTOR ANALYSIS

Cartesian

cylindrical
y

z!=!c3
x!=!c1

x
x

y!=!c2

z!=!c3
r!=!c1

r
!=!c2
x

z
z

spherical
z
r
!=!c
!=!c3

r!=!c1
x

Figure D.3: Orthogonal unit vectors and constant coordinate surfaces for the
three most common orthogonal coordinate systems.

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

23

APPENDIX D. VECTOR ANALYSIS

Note that with these definitions the cylindrical angle is the same as the azimuthal (longitudinal) spherical angle , but that the radial coordinate r is
different in the cylindrical and spherical coordinate systems. The spherical angle is a latitude angle see Fig. D.3. Explicit forms for the various partial
derivatives in space, (D.205) (D.208), are given in Appendix Z.
REFERENCES
Intermediate level discussions of vector analysis are provided in
Greenberg, Advanced Engineering Mathematics, Chapters 13-16 (1998) [?]
Kusse and Westwig, Mathematical Physics (1998) [?]
Danielson, Vectors and Tensors in Engineering and Physics, 2nd Ed. (1997) [?]

More advanced treatments are available in


Arfken, Mathematical Methods for Physicists (??) [?]
Greenberg, Foundations of Applied Mathematics, Chapters 8,9 (1978) [?]
Morse and Feshbach, Methods of Theoretical Physics, Part I, Chapter 1 (1953)
[?]

DRAFT 11:26
October 11, 2002

c
%J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

Appendix F

Transforms, Complex
Analysis
This appendix discusses Fourier and Laplace transforms as they are used in
plasma physics and this book. Also, key properties of complex variable theory
that are needed for understanding and inverting these transforms, and to define
singular integrals that arise in plasma physics, are summarized here.
Fourier and Laplace transforms are useful in solving differential equations because they convert differentiation in the dependent variable into multiplication
by the transform variable. Thus, they convert linear differential equations into
algebraic equations in the transformed variables. In addition, Laplace transforms introduce the (temporal) initial conditions and hence causality into the
transformed equations and the ultimate (inverse transform) solution.

F.1

Fourier Transforms

Fourier transforms are usually used for representing spatial variations because
the spatial domain of the response is often localized away from the boundaries.
For such situations the spatial domain can be considered infinite: |x| . The
Fourier transform F (transformed functions are indicated by hats over them)
and its inverse F 1 are defined in three dimensions by1
!

f (k) = F{f (x)} d3x eikx f (x),


Fourier transform,
(F.1)
!
d3k ikx
ae
f (k), inverse Fourier transform. (F.2)
f (x) = F 1 {f(k)}
e
(2)3
1 The ae above the equal sign in the second equation is there to remind us that the inverse
transform is equal to the original function almost everywhere namely, everywhere the
function f is continuous. At a jump discontinuity the inverse transform is equal to the average
of the function across the discontinuity: [f (x + 0) + f (x 0)]/2.

DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

These three dimensional integrals are defined in cartesian coordinates by


!
!
!
!
!
!
!
!
3
3
dx
dy
dz,
dk
dkx
dky
dkz . (F.3)
dx

Sufficient conditions for the integral in the Fourier transform to converge are that
f (x) be piecewise smooth and that the integral of f (x) converges absolutely:
!
d3x |f (x)| < constant, Fourier transform convergence condition. (F.4)

When these conditions are satisfied, the inverse Fourier transform yields the
original function f (x) at all x except at a discontinuity in the function where it
yields the average of the values of f (x) on the two sides of the discontinuity.
Some useful Fourier transforms are (here k 2 k k)
F{1} =

(2)3 (k),

F{(x x0 )} = e
ik0 x

F{e

} =

F{e|x|/ /|x|} =

(F.5b)

(2) (k k0 ),

(F.5c)

4/(k 2 + 1/2 ),

2 2
F{e|x| /2 } = ( 2 )3 ek /2 ,
F{f (x)} = f(k),
F{f (x)} = i k f(k),

F{ A} = i k A(k),
2

(F.5a)

ikx0

= i k A(k),
F{2 f (x)} = k 2 f(k),
" 3 #

f(k).
F{ d x G(x x# ) f (x# )} = G(k)
F{A(x)}

(F.5d)
(F.5e)
(F.5f)

(F.5)

(F.5g)
(F.5h)
(F.5i)
(F.5j)
(F.5k)

The last relation is called the Fourier convolution relation. Corresponding inverse Fourier transforms can be inferred by taking the inverse Fourier transforms
ae
of these relations and using the fact that F 1 F{f (x)} = f (x).
As can be seen from (F.5e), which is indicative of the Fourier transform
of the smoothest possible localized function in space, the localization in space
(xrms = ) times the localization in k-space (krms = 1/) is subject to the
condition:
(F.6)
k x 1, uncertainty relation.2

Taking the dot product of the Fourier transform of a vector field with its
complex conjugate and integrating over all k-space yields
!
!
d3k
|A(k)|2 , Parsevals theorem.
(F.7)
d3x |A(x)|2 =
(2)3
2 This uncertainty relation indicates the degree of localization in k-space for a given localization of a function in x-space. For the energy density in wave-packets and the probability
density in quantum mechanics, the corresponding uncertainty principle is determined using the
square of the fluctuating field or wave function; then the uncertainty principle is k x 1/2.

DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

F.2

Laplace Transforms

Laplace transforms are often used to analyze the temporal evolution in response
to initial conditions from the present time (t = 0) forward in time, which defines
an infinite half-space time domain (0 < t < ) problem. The Laplace transform
L and its inverse L1 are defined by3
!
f() = L{f (t)}
dt eit f (t),
Laplace transform,
(F.8)
0

f (t) = L1 {f()}
ae

+i

+i

d it
e
f (), inverse Laplace transform. (F.9)
2

Sufficient conditions for the Laplace transform integral to converge are that f (t)
be piecewise smooth and at most of exponential order:
lim f (t) < constant et ,

Laplace transform convergence condition, (F.10)

which defines the convergence parameter needed for the path of integration in
the inverse Laplace transform (F.9). The function f (t) can grow exponentially
in time like et ; then > is required for (F.10) to be satisfied. The obtained
transform f() is only valid for Im{} > . As indicated by the ae (almost
everywhere) over the equal sign in (F.9), the inverse Laplace transform yields
the original function f (t) for all t except at a discontinuity in the function where
it yields the average of the values of f (t) on the two sides of the discontinuity.
Because the original function and its inverse Laplace transform are only valid for
t 0, some people introduce a Heaviside step function H(t) (see Section B.1)
into the integral in the definition of the inverse transform in (F.9) to emphasize
that fact.
3 In

plasma physics it is convenient to use eit as the integrating factor in the definition
of the Laplace transform so that when is real it will represent a (radian) frequency. Many
mathematics texts use est or ept (i s or p) as the integrating factor to emphasize
exponential growth or damping. Most electrical engineering texts use ejt (i j).

DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS


Some useful Laplace transforms are
L{et }

L{eit }

L{et sin(0 t)}

L{et cos(0 t)} =


$
#
2
ex /4Dt

=
L
t
L{H(t)} =
L{(t)}

i
,
+ i
i
,

0
,
( i)2 02
i ( i)
,
( i)2 02

ex i/D

,
i

> ,

(F.11a)

> Im{
}, (F.11b)
> ,

(F.11c)

> ,

(F.11d)

1
i
=
,

i
1,

(F.11e)
(F.11f)
(F.11g)

d(t)
} = i,
L{
dt
% it &
'
e

,
L
=
i(
)
t
L{f (t)} = f(),

(F.11h)
(F.11i)
(F.11j)

L{f(t)} = i f() f (0),


L{f(t)} = 2 f() + if (0) f(0),
1 dn f()
L{tn f (t)} = n
,
i d n
"t #

f().
L{ 0 dt G(t t# ) f (t# )} = G()

(F.11k)
(F.11l)
(F.11m)

(F.11n)
(F.11)
In (F.11b) and (F.11i) the frequency
is in general complex. In (F.11c) and
(F.11d) the frequency 0 and gowth rate are real. In (F.11g) and (F.11h)
the integrals over the delta functions are evaluated by taking account of the
lower limit of the Laplace transform integral being 0 (an infinitesimal negative
time near zero) where the delta function vanishes. The last relation is called
the Laplace convolution relation. Corresponding inverse Laplace transforms
can be inferred by taking the inverse Laplace transforms of these relations and
ae
using the fact that L1 L{f (t)} = f (t). [A Heaviside unit step function H(t)
(see Section B.1) is sometimes inserted to remind one that Laplace transforms
ae
are only defined for t > 0, i.e., L1 L{f (t)} = H(t)f (t).] The simultaneous
localization in time and frequency is subject to a condition similar to (F.6):
t 1,

uncertainty relation.

(F.12)

It is important to be aware of the differences between Fourier and Laplace


transforms. The main difference is that Fourier transforms represent functions
DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

in infinite domains (in space) that have no starting or ending points and no
preferred directions of motion in them. In contrast, Laplace transforms represent functions in an infinite half-space of time that begins (with suitable intitial conditions) at t = 0, increases monotonically, and extends to an infinite
time in the future (t ). These physical differences are manifested mathematically in their transforms of unity. From (F.5a), the Fourier transform of
unity is F{1} = (2)3 (k), which is a function of k that is singular at k = 0.
In contrast, from (F.11a) with 0, the corresponding Laplace transfom is
L{1} = i/, Im{} > 0, which is singular for 0 but with the nature of
the singularity defined (see Sections F.4 to F.6) by the condition Im{} > .
Physically, this condition implies that as time progresses the response grows
less rapidly than et. Thus, Laplace transforms embody the physical property
of causality that the response proceeds sequentially in time from its initial conditions whereas Fourier transforms embody no such directionality in the response
(or dependence on initial or boundary conditions). This key difference is often
highlighted by referring to the relevance of Laplace transforms for initial value
problems and for ensuring temporal causality in the solution.

F.3

Combined Fourier-Laplace Transforms

Often we will need a combination of a three-dimensional Fourier transform in


space and a Laplace transform in time, which is defined by
!
!
dt ei(kxt) f (x, t).
(F.13)
f(k, ) = FL{f (x, t)} d3x
0

The corresponding combined inverse transform is defined by


!
! +i
d3k
d i(kxt)
ae
e
f (k, ). (F.14)
f (x, t) = F 1 L1 {f(k, )}
3
(2) +i 2
For a monochromatic wave [f(k, ) = fk0 ,0 (2)4 (k k0 )( 0 )], we have
f (x, t) = fk0 ,0 ei(k0 x0 t) ,

three-dimensional plane wave.

(F.15)

The representation of f (x, t) in terms of its transform f(k, ) in (F.14) is a


very useful form that is often used (for both scalar functions and vector fields)
and one from which the Fourier and Laplace transforms of spatial and temporal
derivatives in (F.5f)(F.5j) and (F.11j)(F.11l) can be deduced readily.

F.4

Properties of Complex Variables, Functions

A complex variable z = x + iy is a two-dimensional variable (vector) that has


real [x Re{z} zR ] and imaginary [y Im{z} zI ] parts. Its cartesian and
polar angle representations are
(

z = x + iy = zR + izI = rei , r |z| = z z = x2 + y 2 , = arctan y/x.


(F.16)
DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

The function ei is repesented by


ei = cos + i sin , Eulers formula.
(F.17)

Thus, the imaginary unit number i 1 = ei/2 . [More generally, one defines
i = ei(4n+1)/2, n = 0, 1, 2, . . ..] The complex conjugate of z is
z = x iy = |z|ei ,

complex conjugate.

(F.18)

The reciprocal of a complex variable can be written many ways:


1
x iy
x iy
z
ei
1
=
=
= 2
.
= 2 =
2
z
x + iy
(x + iy)(x iy)
x +y
|z|
|z|

(F.19)

A function of a complex variable w(z) wR (z)+i wI (z) is analytic at a point


z zR + i zI if its derivative dw/dz exists there and is the same irrespective
of the direction in the complex z-plane along which the derivative is calculated.
This criterion for a function to be analytic yields the sufficient conditions
wI
wR
=
,
zR
zI

wR
wI
=
,
zI
zR

Cauchy-Riemann conditions for analyticity.

A general expansion of a complex function around z = z0 is


w(z) =

cn (z z0 )n ,

n=

Laurent expansion.

(F.20)

(F.21)

This expansion reduces to a Taylor series expansion if cn = 0 for all n < 0; then,
cn = (1/n!) dnf /dz n |z=z0 , n = 0, 1, 2, . . ..
All functions that are analytic over a region can be expressed in terms of
convergent Taylor series, with the radius of convergence bounded by the distance from the expansion point to the nearest singularity. Examples of (entire)
functions that are analytic over the entire finite z-plane are z, z n, sin z, ez. On
the other hand, the function w1 (z) = 1 + z + z 2 + has a radius of convergence |z| < 1. An analytic function can be analytically continued to adjacent regions where the function is analytic through Taylor series expansion
about other points in the original analytic region or by other means. For example, the power series in the function w1 (z) above can be summed to yield
w1 (z) = 1/(1 z) = 1/(z 1), which can be represented by a Laurent series
with c1 = 1 and z0 = 1 with all other cn = 0. The function 1/(z 1) is
analytic everywhere except at z = 1 and represents the analytic continuation of
the power series respresentation of w1 (z) to all z += 1.
Nonanalytic functions have singularities (z values where they are unbounded
or about which they are multivalued) and are represented by the Laurent series
with cn += 0 for some n < 0. Isolated singularities are classified as follows:
Poles. If the maximum
is m (i.e., cm += 0 and
an mth -order pole at z
first-order pole at z = 1

DRAFT 12:20
August 19, 2003

negative power in the Laurent expansion (F.21)


cn = 0 for n > m), then the function w(z) has
= z0 . For example, w1 (z) = 1/(z 1) has a
and 1/(z 1)2 has a second-order pole at z = 1.

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

Figure F.1: Cauchy integral contours C that: a) do not enclose z0 , b) go


through z0 (really enclose with a small semi-circle), and c) fully enclose z0 .
Essential Singularities. If there are an infinite number of negative powers
present in the Laurent series (F.21), w(z) has an essential singularity at
z0 . For example, e1/z = 1 1/z + 1/2z 2 has an essential singularity
at z = 0 and hence is nonanalytic there. The logarithm function ln z =
ln |z| + i is multivalued (has different values for the same z depending on
which 2 interval is taken to be in) and has an essential singularity at
z = 0 where it is unbounded. Its principal value is usually defined for
0 < 2 with a branch cut inserted at = 2. Additional branches
(Riemann sheets) of ln z are defined for 2 < 4, etc. Since the
encircling of z = 0 is the source ofthe multivaluedness, it is known as
a branch point of ln z. Similarly, z = |z|1/2 ei/2 has a branch point
(essential singularity) at z = 0 and has two branches that are usually
defined for 0 < 2 and 2 < 4.

F.5

Cauchy Integral

The key properties of integration around a simple, closed contour C in the


complex z plane are summarized by a generalized Cauchy integral formula:

!
if C does not enclose z0 , (F.22a)
0,
f (z)
(F.22b)
i f (z0 ), if C goes through z0 ,
(F.22)
dz
=

z z0
C
(F.22c)
2i f (z0 ), if C encloses z0 ,
Cauchy integral formula.

Here, it is assumed that f (z) is an analytic function of z inside and on the


contour C, and motion along the contour is in the counterclockwise direction.
Also, it is assumed for (F.22b) that the contour C goes through the point z0 on
a straight path (i.e., z0 is not at a square corner or other irregular point on C)
and that z0 is on the inside edge of the contour C in a limiting sense. The
contours for the three situations in (F.22) are shown in Fig. F.1.
For a general complex function w(z), (F.22c) generalizes to the residue the-

DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

orem for a contour C that encloses isolated pole-type singularities at z = zj :


!
)
dz w(z) = 2i
c1 (zj ), Cauchy residue theorem.
(F.23)
C

Here, c1 (zj ) is the residue [coefficient c1 in the Laurent expansion (F.21)] of


the function w(z) at the singular point z = zj , which is defined by
c1 (zj ) = lim [(z zj ) w(z)],
zzj

c1 (zj ) =

F.6

first-order pole, (F.24a)

1
d
lim
[(z zj )m w(z)], mth -order pole.
m1
zz
(m 1)!
j dz
m1

(F.24b)
(F.24)

Inverse Laplace Transform Example

To illustrate the use of these complex variable integration formulas (and develop
some inverse transform concepts that are important in plasma physics), consider
their use in evaluating the inverse Laplace transform of the weakly damped
+ x + 02 x = f (x, t). For
( << 0 ) oscillator problem given in (??): x

= 0 [0 = /2 in
simplicity, assume the initial conditions are x(0) = x0 , x(0)
the initial conditions used to derive (??)] and that there is no forcing function f .
Taking the Laplace transform of the homogeneous damped oscillator equation
and solving for the transform of the response, one obtains
S()

x
() = G()
=

x0 ( i)
,
2 i + 02

S()
x0 .

(F.25)

The temporal response x(t) is obtained from the inverse Laplace transform:
! +i
d
I(),
(F.26)
x()} =
x(t) = L1 {
+i 2
I()

eit x0 ( i)
.
( + i/2)( + + i/2)

(F.27)

The integrand I() has first-order poles at = , with residues given by


c1 ( ) =

ei t ( i )
,
2

i/2,

02 2 /4.

(F.28)
Figure F.2a illustrates the inverse Laplace transform integration path (L) in
(F.26) for an arbitrary > 0. As indicated, it is just a line integral from +i
to + i along a line that is parallel to the R Re{} axis, but a distance
I Im{} = above it. While for this problem we could convert this line
integral into a closed contour by adding the (vanishing, for t > 0) integral along
the infinite semi-circle in the lower half -plane [Csc with || as shown
in Fig. F.2a], we will use a more generally useful procedure. [The vanishing of
DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

Figure F.2: Illustration of: a) inverse Laplace transform integration path L and
infinite semi-circle Csc in the lower half -plane which can be used as a closing
contour for t > 0, and b) inverse Laplace transform contour CL and dotted
contour C0 which when added together yield the original integration path L.
the inverse Laplace transform for t < 0 can be shown by closing the contour on
an infinite semi-circle in the upper half plane by observing that because of the
convergence condition (F.10) there are no singularities within this contour.]
For a general Laplace transform inversion procedure, we analytically continue the Laplace integration contour downward, being careful to deform the
contour around the singular points of the integrand, as indicated in Fig. F.2b.
The integral along the original Laplace integration path (L) is equal to the sum
of the Laplace contour CL and the dotted contour C0 between it and the original line integration path (L). However, since there are no singularities of I()
inside the C0 contour, this integral vanishes by (F.22a). Thus, the integral in
(F.26) becomes
!
!
!
! +i
d
d
d
d
I() =
I() +
I() =
I(). (F.29)
2
2
2
+i
C0
CL
CL 2

The CL contour integral includes the two first-order poles at = which are
evaluated4 with (F.24a) using (F.28) for the residues, plus a line integral along
the path i to i which yields a contribution of order et :
!
d
I() = i [c1 (+ ) + c1 ( )] + O{et }
x(t) =
CL 2
/
.

t/2
sin t + O{et }, t 0. (F.30)
= x0 e
cos t +
2
4 The residue integrals are the negative of (F.23) because the small circular contours around
the poles are in the clockwise direction rather than being in the counterclockwise direction
for which (F.22) and (F.23) are defined.

DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

10

The first term is the desired response and is the same as the result (??) obtained
via other means in Section E.2 for the present 0 = /2 case.
The O{et } term in (F.30) represents initial transient responses that decay
exponentially in time for t > 1/. For the present problem since there are no
other singularities in the lower half complex -plane, we can take and
this term vanishes. However, for plasma physics responses there are often many
(sometimes a denumerable infinity of) singularities in the lower half complex plane and we are usually only interested in the time-asymptotic response. Then,
we usually only calculate the responses from the singularities that are highest in
the complex -plane, and estimate the time scale on which this time-asymptotic
response will obtain from the maximum for a contour CL that lies just above
the next highest singularities. Note that the resultant responses may be growing
exponentially in time (if the highest singularities are in the upper half -plane),
and that the transients may also be growing (more slowly) in time (if < 0).
The generic physical points evident from this inverse Laplace transform analysis procedure are that: 1) responses are determined by the singularities of the
integrand of the inverse Laplace transform, which in turn are usually determined
by the singularities of the Laplace transform of the system transfer (Green) func
tion G();
2) the singularities that are highest in the complex -plane dominate
the time-asymptotic response; and 3) the next highest singularities determine
the time scale on which this asymptotic response becomes dominant.

F.7

Ballistic Propagation Example

As another example, we use Fourier-Laplace transforms and complex variable


theory to define the singular responses to ballistic propagation of particles
along straight-line particle trajectories (??): x = x(t = 0) + vt. Consider a simple kinetic equation for a distribution f (x, v, t) with a kinetic source Sf (x, v, t):
f
+ v f = Sf .
t

(F.31)

Taking the Fourier-Laplace transform of this equation using (F.13), (F.5g), and
(F.11k), we obtain
i f f (0) + ik v f = Sf

v, ), (F.32)
f(k, v, ) = G(k,
) S(k,

with transformed source S Sf (k, v, ) + f(k, v, t = 0) in which f represents


just a Fourier transform in space rather than a full Fourier-Laplace transform.

The full transform G(k,


) is in general called a transfer function. Here, it is

G(k,
) =

i
,
kv

Im{} > ,

ballistic propagator.

(F.33)

This Fourier-Laplace transfer function has a singularity at = k v that is


defined (resolved) by the Laplace transform convergence condition (F.10) and
hence by the initial-value problem (causality) characteristics of the Laplace
DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

11

transform. It is called the ballistic propagator in plasma physics because it


represents [in , k transform space see (F.35) below] motion along straightline particle trajectories.
The kinetic distribution f is obtained from the full inverse transform:
!

f (x, v, t) = F 1 L1 {f} =
=

dt#

d3k
(2)3

+i

+i

d i(kxt)
v, )
e
G(k, )S(k,
2

d3x# G(x x#, t t# ) S(x#, v, t# ),

t 0,

(F.34)

in which the second line follows from combining the convolution integrals (F.5k)
and (F.11n) that result from the inverse Fourier and Laplace transforms of the

v, ). The Green function


products of the two transforms G(k,
) and S(k,
G(x, t) is obtained by first using the same inverse Laplace transform procedure
of deforming the Laplace integration contour (see Fig. F.2) downward around
the singularity (in this case at = k v) as was used in the preceding analysis of
the damped oscillator. Then, taking account of the first-order pole, evaluating
the residue via (F.24a), and using the delta function definition in (??) with
x x vt to evaluate the inverse Fourier transform, we obtain
%
& !
d3k ik(xvt)
i
=
e
= (x vt),
G(x, t) = F 1 L1
kv
(2)3
Green function. (F.35)
The inverse Fourier-Laplace transform of S is obtained using (F.2), (F.9) and
L1 of (F.11g):
S(x, v, t) = Sf (x, v, t) + f (x, v, t = 0) (t).

(F.36)

Substituting (F.35) and (F.36) into (F.34), we obtain for t 0


!

d3x# [x x# v(t t# )] S(x# , v, t# )


! t
dt# Sf [x v(t t# ), v, t# ], (F.37)
= f (x vt, v, t = 0) +

f (x, v, t) =

dt#

which is the ballistic response we have been seeking. The first term represents
propagation of the initial distribution function along the ballistic straight-line
particle trajectories x = x(t = 0) + vt, while the second represents the time
integral of the effect of the propagation of the source function along the same
trajectories. Since the solutions propagate (move along) the ballistic motion of
the particles, these are called ballistic solutions. Hence, the transform of the
Green function that caused this response, which is given in (F.33), is called the
ballistic propagator.

DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

12

Figure F.3: Deformation of u integration contour around the singularity (firstorder pole) at u = /k as Im{} decreases from: a) the original definition
region Im{} > > 0, b) to the real axis, and c) to the lower half -plane.

F.8

Singular Integrals In Plasma Physics

Next, we use complex variable theory to define the types of singular integrals
that arise in plasma physics from integrating the ballistic propagator over distribution functions. Defining k v = ku, the types of integrals that arise are of
the form
!
g(u)
du
, Im{} > > 0.
(F.38)
I(/k)
u /k

A sufficient condition
" for this integral to converge is that the integral of g(u) be
bounded (i.e., | du g(u) | < constant). This integral is analytically continued
to lower values of Im{} by deforming the contour around the singularity at
u = /k as Im{} moves from the upper to the lower half -plane, as indicated
in Fig. F.3 for the usual case of k > 0. (An integral in the complex plane
is analytically continued by deforming its integration contour so it is always
on the same side of any pole-type singularities.) Since the integration contour
passes under the singularity for Im{} > 0, through it (but actually on a
small semi-circle below it) for Im{} = 0, and encloses it for Im{} < 0, using
(F.22) we see that I(/k) is defined (for5 k > 0) by
!
g(u)

,
Im{/k} > 0, (F.39a)
du

/k

%!
&
!

g(u)
g(u)
du
+ i g(/k), Im{/k} = 0, (F.39b)
du
P
u /k
u /k

g(u)

+ 2i g(/k), Im{/k} < 0. (F.39c)


du
u /k

(F.39)
5 For k < 0 the integral I(/k) is originally defined for Im{/k} < 0 and analytically
continued to Im{/k} 0, which results in i g(/k) and 2i g(/k) terms (because of
the then clockwise rotation of the integration contour around the pole) on the second and
third lines of this definition which are then applicable for Im{/k} = 0 and Im{/k} > 0.

DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

13

Figure F.4: Areas that cancel in the Cauchy principal value limit process as
) 0 to produce a convergent integral are shown cross-hatched.
For Im{} = 0 the integration over the singularity in the real integrals integrand at u = Re{/k} u0 is defined (i.e., made convergent) by the prescription
&
.! u0 '
/
%!
!
g(u)
g(u)
g(u)
du
lim
du
+
du
,
P
'0
u u0
u u0
u u0

u0 +'
Cauchy principal value operator P. (F.40)
As shown in Fig. F.4, the Cauchy principal value limit process causes the nearly
equal areas on the two sides of the singularity to cancel as ) 0; it thereby
yields a finite integral as long as g(u) is a continuous function of u at u = u0 .
The definition of I(/k) in (F.39) appears to be discontinuous as Im{}
approaches zero from above and below, but is in fact continuous there. In the
limit of Im{} ) 0, the singular part of the integrand becomes
%
&
1
(u u0 ) i)
1
= lim
=
P
i (u u0 ),
lim
'0 u (u0 i))
'0 (u u0 )2 + )2
u u0
Plemelj formulas.(F.41)
In obtaining the last, imaginary term, we used the definition of the delta function from (??) and (??) in Section B.2. Using the Plemelj formulas, it can be
shown that the Im{} 0 limits of both (F.39a) and (F.39c) yield (F.39b).
Thus, the definition in (F.39) is just what is needed to make I(/k) a continuous
function of Im{}; hence, (F.39) represents the proper analytic continuation of
the function I(/k) defined in (F.38) from the upper half -plane, where it
is initially defined, to the entire -plane. Note also that since the representations in the various Im{} regions are continuous in the vicinity of Im{} . 0,
we can use any of the representations there. In plasma physics the representation of I{/k} for Im{} = 0 given in (F.39b) is often used for all Im{} . 0.
DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX F. TRANSFORMS, COMPLEX ANALYSIS

14

REFERENCES
Discussions of transforms and complex variable theory are provided in most advanced engineering mathematics and mathematical physics textbooks, for example:
Greenberg, Advanced Engineering Mathematics (1988,1998), Chapts. 5, 2124
[?]
Greenberg, Foundations of Applied Mathematics (1978), Chapts. 6, 1116 [?]
Morse and Feshbach, Methods of Theoretical Physics (1953), Vol. I, Chapt. 4
[?]
Arfken, Mathematical Methods for Physicists (1970) [?]
Kusse and Westwig, Mathematical Physics (1998), Chapts. 69 [?]
Classic treatises on the theory of complex variables are
Whittaker and Watson, A Course of Modern Analysis (1902,1963) [?]
Copson, Theory of Functions of a Complex Variable (1935) [?]
Carrier, Crook, Pearson, Functions of a Complex Variable (1966) [?]
An extensive table of Fourier and Laplace (and other) transforms is provided in
Erd`elyi, Tables of Integral Transforms, Vol. 1 (1954) [?]

DRAFT 12:20
August 19, 2003

c
$J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX Z. USEFUL FORMULAS

Appendix Z

Useful Formulas

DRAFT 11:19
February 2, 2002

c
!J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX Z. USEFUL FORMULAS

Key Vector Relations


AB

A (BC)

A(BC)

(AB) (CD)

(AB)(CD)

B A,

AB

(AB) C

BA,

B (A C) C (A B),

AA

0,

bac-cab rule

(A C) (B D) (A D) (B C)

C (AB D) D (AB C)

+ A with b
B/B
A = A! b
A
A! B A/B = b

bA)

A B(BA)/B 2 = b(

A = (B ) (A! /B) + (A! /B) ( B) + A

)b
] (1/B) b
(BA )
A = A [ ln B + (b

For A = Bf /B 2 ,
(f g)

(f A)

=
=

(f A)

(f T)

(f T)

(BA ) = (b
f ) (b
b)

gf + f g

f A + f A

f A + f A

f
A

2 A

f T + f T

f T + f T

(B )(A C)

(AB)

(AB)

(A B)

(AB)

= ( A) (A)

C (B ) A + A (B ) C

A(B) + B(A) + (A ) B + (B ) A
B ( A) + (A ) B
B A A B

A ( B) B ( A) + (B ) A (A ) B

ey + z
ez and |x|
For the general coordinate x x
ex + y
x = 3,

2 f

x = 0,

x = I,

I = 0,

!
x2 + y 2 + z 2 ,

I = 0,

|x| = x/|x|, (1/|x|) = x/|x| , (1/|x|) = 4(x),


3

A I = A,

I A = A.

For a volume V enclosed by a closed, continuous surface S,


""
" 3
d x A = !S dS A, divergence, Gauss theorem.
V

For an open surface S bounded by a closed, continuous contour C,


#
""
dS A = C d! A, Stokes theorem.
S
DRAFT 11:19
February 2, 2002

c
!J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX Z. USEFUL FORMULAS

Explicit Forms Of Vector Differentiation Operators


(
ei ui /|ui |, Ai
ei A)

Cartesian coordinates: ui = (x, y, z),


f

2 f

"

d3x =

"

dx

"

dy

"

dz,

f
f
f
+
ey
+
ez
x
y
z
Ay
Az
Ax
+
+
x
y
z
%
%
%
$
$
$
Ay
Az
Ax
Az
Ax
Ay

+
ey

+
ez

ex
y
z
z
x
x
y

ex

2f
2f
2f
+ 2 + 2
2
x
y
z

" 3
"
" 2
"
i
Cylindrical
coordinates:
u
=
(r,
,
z),
d
x
=
r
dr
d
dz,
0
0

!
2
2
with r x + y , arctan (y/x), z z,
and inverse relations x = r sin , y = r cos , z = z,
f

2 f

f
1 f
f
+
e
+
ez
r
r
z
Az
1
1 A
(rAr ) +
+
r r
r %
z
$
$
%
A
Az
1 Az
Ar

er

+
e

r
z
z
r
$
%
Ar
1
(rA )
+
ez
r r

&
'
1
f
1 2f
2f
r
+ 2 2 + 2
r r
r
r
z

er

" 3
" 2
"
" 2
i
Spherical
! d x = 0 r dr 0 d sin 0 d,
!coordinates: u = (r, , ),
with r x2 + y 2 + z 2 , arctan ( x2 + y 2 /r), arctan (y/x),
and inverse relations x = r sin cos , y = r sin sin , z = r cos ,
f

2 f

f
1 f
1 f
+
e
+
e
r
r
r sin

1 2
1
1 A
(r Ar ) +
(sin A ) +
2
r r
r sin
r sin
$
%

1
A
(sin A )

er
r sin

%
$
%
$
1
1
Ar
1 Ar

(rA ) +
(rA )
e
+
e
r sin
r r
r r

&
'
&
'
1

f
1
2f
1
f
r2
+ 2
sin
+ 2 2
2
r r
r
r sin

r sin 2

er

DRAFT 11:19
February 2, 2002

c
!J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX Z. USEFUL FORMULAS

Physical Constants
me
mp
mp /me
e
c
0
'0
h
NA
e/kB

electron mass
proton mass
mass ratio
elementary charge
speed of light in vacuum
permeability of vacuum
permittivity of vacuum
Planck constant
Avogadro constant
Boltzmann constant

9.11 1031 kg, 511 keV


1.67 1027 kg, 938 MeV
1836 = (42.85)2
1.602 1019 C (= J/eV)

3 108 m/s = 1/ 0 '0


4 107 N/A2
8.85 1012 F/m, 4'0 ' 1010
6.626 1034 J s
6.022 1023 #/mol
11 600 K/eV

Key Plasma Formulas


Quantities are in SI (mks) units except temperature and energy which are expressed in eV; Zi is the ion charge state; Ai mi /mp is the atomic mass number.
Frequencies
electron plasma

pe

ion gyrofrequency

ci

electron collision

Lengths
electron Debye

De

ion gyroradius

+i

electron collision

Speeds, Velocities
electron thermal

vT e

ion thermal

vT i

ion acoustic (Te >> Ti ) cS

qi B
Zi B
' 0.96 108
rad/s
mi
Ai
&
'
4
5 1011 ne Zi ln
(vT e ) '
s1
17
3
[Te (eV)]3/2

(
'0 Te
Te (eV)
3
' 7.4 10
m
ne e2
ne
!
Ti (eV) Ai
vT i
m
' 1.4 104
ci
Zi B
'
&
2
vT e
17
16 [Te (eV)]
m
' 1.2 10
e
ne Zi
ln

Alfven
electron diamagnetic
flow (dTe /dx = 0)

Ve

electron drift in B(x)


(average, low )

De
v

DRAFT 11:19
February 2, 2002

cA

ne e2
me '0 ' 56 ne rad/s, fpe ' 9 ne Hz

!
!
2 Te /me ' 5.9 105 Te (eV) m/s
!
!
2 Ti /mi ' 1.4 104 Ti (eV)/Ai m/s
!
!
Zi Te /mi ' 104 Zi Te (eV)/Ai m/s

B/ 0 m ' 2.2 1016 B/ ni Ai m/s


&
'
1 dne
Te
Te (eV)

ey =

ey m/s
qe B ne dx
B Ln
&
'
2Te 1 dB
2 Te (eV)

ey =

ey m/s
qe B B dx
B LB

c
!J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX Z. USEFUL FORMULAS


Drift, flow velocities (for + << 1, << c ) perpendicular to B :

plasma species flow velocities

particle drift velocities


vF = FB/qB 2
vE = EB/B 2
v = BB/qB 2
2
/2B
mv
v = Bmv!2 /qB 2
)b
= RC /R2
(b

vp = Bm(dvD /dt)/qB 2

vD = vE + v + v

general force
EB
grad-B

VF = FB/qB
2
VE = EB/B

curvature
diamagnetic
polarization
friction
viscosity
total

V
Vp
V
V

21 [Te (eV)]

no magnetic field

e 2e

vT2 e /e

magnetic field

/0

(me e /ne e2 )/
& 0

e +2e

Bp/nqB 2
Bm(dV/dt)/qB 2
RB/nqB 2
B /nqB 2

V = VE + V + V p + V + V

Diffusivities

classical

=
=
=
=

' 7 10

ne Zi

5/2&

'&

' 1.4 10

Zi
[Te (eV)]3/2

' 5.6 10

ne Zi
B 2 [Te (eV)]1/2

e (/0 )

22

Dimensionless
number of electrons
in Debye cube

ne 3De

'

Coulomb logarithm

ln

17
ln

ln
17
&

'

ln
17

'

m2 /s

m2 /s
'

m2 /s

4.1 1011 [Te (eV)]3/2 / ne


&
'
D
ln
qm
max [ bcl
min , bmin ]

2
10
bcl
Zi /Te (eV) m
min = Zi /(12ne De ) ' 5 10

plasma to
magnetic pressure

Lundquist number

DRAFT 11:19
February 2, 2002

10
/[Te (eV)]1/2 m
bqm
min = h/(4me v) ' 1.1 10
n e T e + ni Ti

B 2 /20
%
) *$
ni
25 ne
' 4.0 10
Te (eV) +
Ti (eV)
B2
ne
2
a /(/0 )
S
L! /cA
'
&
2
(eV)]3/2
17
13 a B [Te
' 1.6 10
ln
L! Zi ni Ai

c
!J.D
Callen,

Fundamentals of Plasma Physics

APPENDIX Z. USEFUL FORMULAS

Fundamental Equations of Physics


Mechanics
ma
F

dp/dt

m dv/dt = F, v dx/dt

Newtons second law

q (E + vB)

Lorentz force

|p qA|2 /2m + q , p = mv + qA

Hamiltonian

H/q, dq/dt = H/p

Hamiltons equations

Electrodynamics
=

q /'0

Gausss law

B/t

Faradays law

no magnetic monopoles

0 (J + '0 E/t)

Amperes law, 0 '0 = 1/c2

charge continuity equation

q /t + J

E
B

A/t, B = A

potential representations

Plasma Physics
Plasma kinetic equation (PKE) for distribution function f fs (x, v, t):

f / t + v f / x + (q/m) (E + vB) f / v = C(f ).

Density, flow moments and charge, current densities:


"
"
+
+
ns d3v fs , Vs d3v v fs /ns , q s ns qs , J s ns qs Vs .

Gibbs (A: adiabatic) distribution of plasma species with temperature T:


)
*3/2
m
eH/T ; nA (x, t) = n0 eq/T , Boltzmann relation.
fA = n0 2T
!
Maxwellian (collisional equilibrium) distribution (vT 2T /m):
&
'
2
)
*3/2
2
v "2 /vT
m|v& |
m
exp 2T
= n e 3/2 3 , v& v V.
fM = n 2T
vT
Species fluid moment equations (density, momentum, energy):
"
2
n/t + nV = 0,
nT d3v (mv & /3) f ,
mn (dV/dt) = nq (E + VB) p + R,

(3/2)(n dT /dt) + p V = q : V + Q,

d/dt /t + V ,
p nT .

Magnetohydrodynamics (plasma fluid description,


isotropic
+
+ pressure and
isentropic responses for plasma species, m s ns ms , V s ns ms Vs /m ):
m /t + m V = 0,

m (dV/dt) = JB P,

DRAFT 11:19
February 2, 2002

E + VB = J,
,
d ln P/m /dt = 0,

c
!J.D
Callen,

ps .

Fundamentals of Plasma Physics

You might also like