Sie sind auf Seite 1von 160

Plasma Physics

Don Melrose
and Alex Samarian
Senior-level (3rd year) course
Lecture notes
Version: April 4, 2011
ii
Preface
This course was given for the rst time in 2009, and it has been revised and re-arranged for 2011.
Objectives in this course are
(i) to provide an understanding the physics of fundamental phenomena in plasmas, and
(ii) to familiarize the students with the basic methods of theoretical and experimental plasma
physics.
The class will be separated into Normal and Advanced streams for the nal six lectures.
Chapters 1013 in these notes are for the Advanced class only. (Lecture 14 will be included if
time permits.) Alex Samarian will give the six lectures for the Normal class; the notes for these
lectures will be provided separately.
I shall be absent from Sydney for the rst two lectures (14 and 15 April 2011) and for three
lectures in the rst week of May (after the mid-semester break). These lecture will be given by
Joe Khachan.
Don Melrose
March, 2011
iv
Contents
Preface iii
1 Plasma: an ionized gas 1
1.1 Arc discharges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Ionosphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Solar atmosphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Controlled fusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 Solar radio astronomy and space plasma physics . . . . . . . . . . . . . . . . . . . 7
1.6 Numerical values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.7 Conversion factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.8 Exercise Set 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2 Collective eects in plasmas 13
2.1 Electrostatic oscillations in cold plasma . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Fluid model for electron gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Collective response to a static eld . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4 Debye screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5 Plasma parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.6 Exercise Set 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Exercise Set 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3 Wave dispersion in plasmas 21
3.1 Sound waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2 Expansion in plane waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
vi CONTENTS
3.3 Phase and group velocities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.4 Exercise Set 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4 Waves in isotropic plasmas 29
4.1 Wave equation for a plasma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2 Wave equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.3 Longitudinal response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.4 Cold electron gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.5 Energetics in waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.6 Exercise Set 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5 Orbit theory 39
5.1 Motion of a charged particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.2 Electric drift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.3 Drift motions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.4 Adiabatic invariant: magnetic trapping . . . . . . . . . . . . . . . . . . . . . . . . 45
5.5 Appendix: formal theory of adiabatic invariants . . . . . . . . . . . . . . . . . . . 46
5.6 Exercise Set 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6 Cold magnetized plasma 51
6.1 Response of a cold plasma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.2 Dispersion equation for a cold plasma . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.3 Polarization vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.4 Polarization of cold plasma waves . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.5 Exercise Set 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
7 Magnetoionic theory 59
7.1 Magnetoionic parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
7.2 Cuto frequencies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
7.3 High-frequency limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
7.4 Polarization of magnetoionic waves . . . . . . . . . . . . . . . . . . . . . . . . . . 63
7.5 Exercise Set 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
CONTENTS vii
8 Magnetohydrodynamics 69
8.1 MHD equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
8.2 Ideal MHD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
8.3 Energy density and energy ow . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
8.4 Magnetic pressure and magnetic tension . . . . . . . . . . . . . . . . . . . . . . . 74
8.5 MHD equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
8.6 Exercise Set 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
9 MHD waves 81
9.1 Linearized MHD equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
9.2 Alfven mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
9.3 Magnetoacoustic waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
9.4 Fast and slow modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
9.5 Exercise Set 9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
10 Collisions (Coulomb interactions) 87
10.1 Coulomb interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
10.2 Straight-line approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
10.3 Collision frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
10.4 Electrical conductivity and runaway electrons . . . . . . . . . . . . . . . . . . . . 91
10.5 Damping of waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
10.6 Exercise Set 10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
11 Landau damping 95
11.1 Landau damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
11.2 Dispersion and dissipation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
11.3 Negative Landau damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
11.4 Exercise Set 11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
12 Kinetic theory 103
12.1 Phase space and distribution functions . . . . . . . . . . . . . . . . . . . . . . . . 103
12.2 Boltzmann and Vlasov equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
12.3 Linearized Vlasov equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
viii CONTENTS
12.4 Tensor notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
12.5 Isotropic plasma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
12.6 Thermal plasma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
12.7 Exercise Set 12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
13 Isotropic thermal plasma 111
13.1 Plasma dispersion function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
13.2 Properties of Langmuir waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
13.3 Ion sound waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
13.4 Debye screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
13.5 Exercise Set 13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
14 Plasma instabilities 119
14.1 Reactive beam instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
14.2 Bump-in-tail instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
14.3 Current-driven instabilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
14.4 Exercise Set 14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
A Cartesian tensors 127
A.1 Vector components and rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
A.2 Rotation matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
A.3 Tensor equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
A.4 Rules for tensor equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
A.5 Vector calculus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
B Fourier transforms 135
B.1 Fourier integral theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
B.2 Properties of Fourier Transforms . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
B.3 The Dirac -function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
B.4 Truncations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
B.5 Relations involving generalized functions . . . . . . . . . . . . . . . . . . . . . . . 141
B.6 Alternative treatment of Plemelj formula . . . . . . . . . . . . . . . . . . . . . . . 144
CONTENTS ix
C Assignment Sets 145
Assignment Sets 145
C.1 Assignment Set 1: Due 20 May 2011 . . . . . . . . . . . . . . . . . . . . . . . . . 145
C.2 Assignment Set 2: Due 3 June 2011 . . . . . . . . . . . . . . . . . . . . . . . . . . 147
x CONTENTS
Chapter 1
Plasma: an ionized gas
We are familiar with three states of matter: solid, liquid and gas. Plasma is sometimes regarded
as a fourth state of matter. It is interesting that the ancient greeks considered there to be four
elements: earth, water, air and re. The rst three may be interpreted as solid, liquid and gas,
and re is one example of a plasma. In re there is a chemical reaction (usually oxidization) in
which atoms tare briey ionized, emitting optical radiation as they recombine. It is the presence
of electrons and charged ions that characterizes the plasma state.
The name plasma for an ionized gas was introduced by Langmuir
1
in the late 1920s in connec-
tion with an investigation of oscillations in an arc discharge. (I am not aware of the reason for the
name plasma.) A plasma may be dened as an ionized gas. There are also solid-state plasmas
such as metals, where the electrons in the conduction band can be regarded as an electron gas for
many purposes. In fact most of the matter in the Universe is ionized, with our environment on
the surface of the Earth being exceptional. Above us, the ionosphere is ionized, and below us, the
Earths core is a strongly conducting solid-state plasma.
In this rst lecture several examples of plasmas, and the physics of relevance to them, are
discussed, following a roughly historical sequence in the development of plasma physics as a
separate branch of physics.
1
L. Tonks, I. Langmuir, Oscillations in ionized gases, Physical Review 33, 195 (1929)
2 1. Plasma: an ionized gas
1.1 Arc discharges
The rst research work on what is now recognized as plasma physics was that of Langmuir in
the late 1920s on oscillations in a gas ionized by an electric arc discharge. In Langmuirs experi-
ments, the plasma was created by a large electric eld driving a large current through the plasma,
such that electrons accelerated by the electric eld knock further electrons o atoms, creating a
relatively high density of electrons and positive ions. Langmuir observed oscillations in the arc
discharge at a characteristic frequency, called the plasma frequency,
p
, that is proportional to the
square root of the number density of electrons, n
e
. Dened as an angular frequency, the plasma
frequency is determined by

p
=
_
e
2
n
e

0
m
e
_
1/2
= 56.4
_
n
e
1 m
3
_
1/2
s
1
, (1.1)
where e and m
e
are the charge and mass of the electron, respectively. The plasma-electron oscil-
lations observed by Langmuir are now called Langmuir waves. Langmuir waves have frequencies
close to
p
, and slightly above it. Langmuir waves are longitudinal or electrostatic: they have
an electric eld along the direction of wave propagation, and no magnetic eld. This makes them
quite dierent from electromagnetic waves, which are referred to as transverse waves in plasma
physics.
Transverse waves are similar to electromagnetic waves in vacuo, in the sense that the electric
eld, E, in the wave is orthogonal to the direction of wave propagation. The refractive index in a
plasma is less than unity: n = (1
2
p
/
2
)
1/2
, with the (angular) frequency of the wave.
Langmuir actually identied two new types of wave, with the other type now being called ion
acoustic waves or ion sound waves. They are also longitudinal. Ion acoustic waves exist only
below the ion plasma frequency,
pi
, which is dened similar to (1.1) with the charge and mass
replaced by those of the ions:
pi
= (Z
2
i
e
2
n
i
/
0
m
i
)
1/2
, where the ions are assumed to have charge
Z
i
e, mass m
i
and number density, n
i
.
Note that a frequency can be expressed either as an angular frequency, , of as cyclic a
frequency, f = /2. The units of are radians per second (s
1
), and the units of f are hertz
(Hz), or cycles per second in older literature.
A thermal distribution is described by its temperature. In a plasma the temperature, T
e
, of
the electrons can be dierent from the temperature, T
i
, of the ions. There are various heating
1.2 Ionosphere 3
processes for plasmas, and most of these favor either the electrons or the ions. Collisions tend to
bring the electron and ion temperatures into equilibrium, but often the time required for T
e
to
approach T
i
is much longer than the heating time. More generally, collisional processes can be
quite slow in plasmas, especially for fast electrons and ions. A simple model for a plasma consists
of thermal electrons and thermal ions, with T
e
,= T
i
, and various distributions of nonthermal
fast particles. Collisions can have a negligible eect on the fast particles, and the physics of
collisionless plasmas is dominated by interactions of waves and particles.
In plasma physics one often uses energy units to describe the temperature, omitting Botzmanns
constant. The temperate of an arc discharge is typically measured in electron volts (eV), with the
ionization potential for most ions being several eV. The conversion factor between kelvin (K) and
eV is given in Table 1.4. It is convenient to dene the thermal speed, V
e
, of electrons by writing
m
e
V
2
e
= T
e
(note: no factor of 1/2 is included). In Langmuirs theory, T
e
appears through the
Debye length:

D
=
V
e

p
= 69
_
T
e
1 K
_
1/2
_
n
e
1 m
3
_
1/2
m. (1.2)
In modern day terminology, the plasma investigated by Langmuir is regarded as an unmag-
netized thermal plasma. Most plasmas of interest are either conned by a magnetic eld, or are
threaded by a magnetic eld that is frozen into the plasma, and such plasmas are said to be
magnetized.
1.2 Ionosphere
The Earths atmosphere is ionized by ultraviolet radiation from the Sun, and by cosmic rays. The
degree of ionization is determined by a balance between the ionization rate and the recombination
rate. The recombination rate is strongly dependent on density, and is very high near the surface of
the Earth, so that the degree of ionization is extremely small. The density decreases rapidly with
increasing height, and around a height of 100 km, the recombination rate becomes comparable
with the ionization rate. Above this height, the degree of ionization is relatively high and this
region is called the ionosphere. Initially, the electron density increases with height, due to the
rapidly increasing degree of ionization, until the plasma becomes completely ionized, and at still
greater height, the electron density decreases as the total density decreases.
The maximum plasma frequency in the ionosphere is where the electron density is a maximum.
4 1. Plasma: an ionized gas
This maximum is several megahertz, depending on the latitude, the time of day and the season.
This frequency is in the radio range; it is above the frequencies used for AM radio transmission
and below the frequencies used for FM radio and TV transmissions. Radio waves in a plasma are
transverse waves, which cannot propagate below the plasma frequency. AM radio waves reect
o the ionosphere and are trapped below it. This allows one to receive a signal from an AM
radio station that is a relatively long way away; the signal can arrive after many reections o
the ionosphere and o the surface of the Earth. However, FM, TV and UHF signals propagate
through the ionosphere and escape. To get good FM, TV and UHF reception one needs to have
a direct line of sight to the transmitter.
The Earths magnetic eld aects the motion of electrons in the ionosphere, causing them to
gyrate about magnetic eld lines at the electron cyclotron frequency,

e
=
eB
m
e
= 1.76 10
11
_
B
1 T
_
s
1
. (1.3)
Magnetic elds are often expressed in terms of an older unit called a gauss (G), with 1 G = 10
4
T,
so that (1.3) becomes
e
= 1.7610
7
(B/1 G) s
1
. The Earths magnetic eld implies a cyclotron
frequency that is also in the radio range, about 0.5 MHz near the magnetic poles, and half that
value near the equator.
The theory for propagation of radio wave in the ionosphere was developed in the early 1930s,
and is called the magnetoionic
2
theory. In this theory the eect of the Earths magnetic eld
is taken into account in the cyclotron motion of electrons. The electrons are regarded as cold,
implying that their thermal motions are neglected (T
e
0). The ions are ignored, except that
the collision frequency between ions and electrons is sometimes included. In modern terminology,
the magnetoionic theory corresponds to wave propagation in a cold magnetized electron gas.
The inclusion of the magnetic eld causes radio wave propagation in the ionosphere to be
anisotropic, which implies that the refractive index depends on the angle of propagation, , relative
to direction of the magnetic eld. The transverse waves of an unmagnetized plasma separate
into two dierent modes, called the ordinary (o) and extraordinary (x) modes, by analogy with
propagation of light in an anisotropic crystal. Like transverse waves in an unmagnetized plasma,
the o- propagates only above the plasma frequency, >
p
, with its refractive index becoming
zero at =
p
. The and x-modes has a smaller refractive index than the o mode, and it becomes
2
Magnetoionic is based on a now-outdated use of ion to mean any ionized particle, including electrons.
1.3 Solar atmosphere 5
zero (called a cuto) at a frequency slightly greater than the plasma frequency.
The properties of these wave are important in understanding ionospheric sounding. In iono-
spheric sounding, the properties of the ionosphere are inferred by studying how radio waves emitted
from the transmitter reect and are detected at a receiver some distance from the transmitter.
The radio wave separates into the two modes, which propagate along slightly dierent ray paths,
and reect at dierent heights. The x mode reects o the ionosphere at a slightly lower height
than the o mode. The two modes have opposite circular polarizations when the return to the
Earth, and they produce dierent signals (the o trace and the x trace) on a chart recorder.
There are two other modes in the magnetoionic theory that propagate at lower frequencies:
the z-mode and the whistler mode. The magnetoionic theory is described in more detail in a later
lecture.
1.3 Solar atmosphere
The atmosphere of Sun can be separated into three regions: the visible surface is called the
photosphere, a thin layer above this is called the chromosphere, and the region above this, out to
several solar radii, is called the corona. The solar corona joins on continuously to the solar wind,
which ows past the Earth and to the edge of the so-called heliosphere at several hundred AU.
3
The temperate in the photosphere is about 10
4
K, decreasing slowly with height before increasingly
abruptly, at the so-called transition region, to > 10
6
K in the corona. The Sun has a magnetic eld,
which has a dipolar component of about 10
4
T, which reverses about every 11 years in the solar
cycle. There are much stronger magnetic elds, B 10
1
T near the photosphere, associated with
magnetic ux loops that emerge from below the photosphere, and cause solar activity, including
solar ares. The heating of the solar corona and solar ares involve a variety of plasma physics
processes, not all of which are yet understood.
The study of the properties of the magnetized plasma in the solar atmosphere led to a uid
description of the plasma, rst developed by Alfven in the early 1940s. In a uid theory, the
important parameters are the uid density, described by the mass per unit volume, say, and the
uid velocity, u say. Fluid theory for an un-ionized gas is referred to as hydrodynamics. There
are two important changes when the gas is ionized. One is that there is an additional force on
3
1 AU is the distance between the Sun and the Earth
6 1. Plasma: an ionized gas
the uid. In an un-ionized gas, besides mechanical forces such as gravity, the important force
is a pressure gradient. In the presence of a magnetic eld there is an additional force (per unit
volume) J B, where J is the current density. Using curl B =
0
J, this force may be interpreted
as a magnetic pressure gradient plus a tension along the magnetic eld lines. The other important
change is that the electrical conductivity, , is very high, whereas it is zero in an un-ionized gas.
In the rest frame of the uid, the current density and the electric eld are related by J = E, and
for , to avoid an innite current density, one must have E = 0. If the uid has a nonzero
velocity, the electric eld in its rest frame is E +u B and requires
E+u B = 0. (1.4)
An important implication of the condition (1.4) is that it means that the magnetic eld lines are
frozen into the uid, in the sense that they move at the uid velocity, u. The generalization of
hydrodynamics to include these eect is called magnetohydrodynamics (MHD). MHD is discussed
in more detail in a later literature.
Alfven recognized that the tension associated with magnetic eld lines cause them to act like
stretched strings, allowing waves to propagate just like the waves on an elastic string. These waves
are now called Alfven waves. They propagate at the Alfven speed,
v
A
=
B
(
0
)
1/2
= 2.2 10
16
_
B
1 T
__
n
e
1 m
3
_
1/2
ms
1
, (1.5)
where the plasma is assumed to be ionized hydrogen.
1.4 Controlled fusion
Starting circa 1950, and continuing today, an important application of plasma physics is to con-
trolled thermonuclear fusion. The objective is create a suciently hot and dense plasma that
nuclei (e.g., two deuterons) fuse together, releasing the binding energy in a controlled version of
the energy release in a hydrogen bomb. It is necessary to conne this hot plasma for long enough
4
for the energy released to be collected and used to generate electricity. A useful fusion machine
requires that the energy generated exceed the energy used in creating the plasma. Many dierent
aspect of plasma physics need to be understood in connection with fusion research. These include
4
Quantied in the Lawson criterion, see Wikipedia for details.
1.5 Solar radio astronomy and space plasma physics 7
orbit theory, needed to understand the trapping of particles in magnetic elds, various plasma
heating mechanisms, which are essential to achieving the very high temperatures needed, various
diagnostic techniques, needed to determine the temperature and density of the plasma without
being able to insert probes into it, instability theory, needed to avoid instabilities that would
disrupt the magnetic connement of the plasma, and radiation processes, that can be important
energetically and diagnostically. Some of these topics will be discussed in detail in later lectures.
From a historical perspective, the interest in controlled fusion led to research on plasma physics
being classied in the early 1950s. This led to plasma theory developing in signicantly dierent
ways in Russia and in the West. The inuence of these dierent approaches remained long after
the work was declassied, and published in 1956.
1.5 Solar radio astronomy and space plasma physics
After the end of World War II the expertise that had been developed in radar physics was redirected
and led to the rapid development of the new eld of radio astronomy, in which Australia became
a world leader. Most of the early observations were at frequencies from tens to hundreds of
megahertz. The Sun emits bursts of radio waves, with prominent types (called types I, II and
III) of solar bursts involving emission at the plasma frequency or its second harmonic. The radio
emission is generated in a two-stage process. The rst stage is generation of Langmuir waves, with
the main mechanism being an instability, which for type III bursts is due to a stream or beam of
fast electrons propagating outward through the corona. The second stage is conversion of these
Langmuir waves into transverse waves at the fundamental and second harmonic of the plasma
frequency through nonlinear processes in the plasma. This is called plasma emission.
Starting around 1970, observations at much lower frequencies from spacecraft became possible.
It was found that the electron streams that generate type III bursts propagate through the solar
wind to the Earths orbit (at 1 AU, where AU stands for astronomical unit) and beyond. Plasma
emission was also observed from the bow shock, where the solar wind hits the Earths magnetic
elds, and from other shock embedded in the solar wind. This led to the development of a eld
that is now called space plasma physics. Spacecraft have no explored out to about 100 AU, and the
plasma emission has been observed from the termination shock where the solar wind is stopped
by the interstellar medium.
8 1. Plasma: an ionized gas
1.6 Numerical values
Relevant physical constants and plasma constants, are listed in Table 1.1 and Table 1.2, respec-
tively, Values are given in both SI and gaussian (cgs) units.
Table 1: Physical Constants
physical quantity symbol SI units gaussian units
speed of light c 3.0 10
8
ms
1
3.0 10
10
cms
1
fundamental charge e 1.6 10
19
C 4.8 10
10
esu
electron mass m
e
9.1 10
31
kg 9.1 10
28
g
proton mass m
p
1.67 10
27
kg 1.67 10
24
g
mass ratio m
p
/m
e
1837 1837
Boltzmann constant
B
1.38 10
23
J K
1
1.38 10
16
erg K
1
electron volt eV 1.6 10
19
J 1.6 10
12
erg
classical e

radius r
0
2.8 10
15
m 2.8 10
13
cm
Thomson cross section
T
6.65 10
29
m
2
6.65 10
25
cm
2
permittivity of free space
0
8.85 10
12
Fm
1
permeability of free space
0
1.23 10
6
Hm
1
1.7 Conversion factors 9
Table 2: Plasma physics quantities
quantity symbol SI units gaussian units
plasma frequency
p
56.4n
1/2
e
s
1
5.64 10
4
n
1/2
e
s
1
f
p
=
p
/2 9.0n
1/2
e
Hz 9.0 10
3
n
1/2
e
Hz
electron gyrofrequency
e
1.76 10
11
Bs
1
1.76 10
7
Bs
1
f
B
=
e
/2 2.8 10
10
BHz 2.8 10
6
BHz
proton gyrofrequency
p
0.98 10
7
Bs
1
0.98 10
4
Bs
1
Alfven speed v
A
2.2 10
16
B(n
e
)
1/2
ms
1
2.2 10
11
B(n
e
)
1/2
cms
1
sound speed c
s
1.5 10
2
T
1/2
ms
1
1.5 10
4
T
1/2
cms
1
ion sound speed v
s
9 T
1/2
e
ms
1
9.1 10
3
T
1/2
e
cms
1
Debye length
D
69T
1/2
e
n
1/2
e
m 6.9T
1/2
e
n
1/2
e
cm
thermal e

speed V
e
3.9 10
3
T
1/2
e
ms
1
3.9 10
5
T
1/2
e
cms
1
collision frequency
0
1.37 10
5
(ln)n
e
T
3/2
e
s
1
13.7(ln)n
e
T
3/2
e
s
1
where is the number of nucleons per electron, = 1 for a hydrogen plasma.
1.7 Conversion factors
Conversion factors between dierent systems of units are made most conveniently by introducing
conversion factors and regarding units as algebraic symbols. For example, consider conversion from
meters to centimeters, or vice versa. Given 10
2
cm = 1 m, one may introduce conversion factors by
writing 1 = 10
2
cmm
1
or 1 = 10
2
mcm
1
, both of which follow directly from the basic relation.
Then if one is given a formula in meters, as L = xm, where x is a number, the quantity in
centimeters, that is, L = y cm where y is another number, is given by L = xm(10
2
cmm
1
) =
y cm, so that one identies y = 10
2
x.
The detailed formulae in these notes are given in SI units. Another set of unit used widely
are gaussian units, also called cgs units because the basic units of length, mass and time are
centimeter, gram and second, whereas SI units are mks units (meter, kilogram second). The
electric and magnetic units are also dierent in SI and gaussian units, and this results in formulae,
including Maxwells equations, having dierent forms in the two sets of units. To convert any
formula from SI to gaussian units involves no change to the charge q, charge density , current
density J, electric eld E and electric potential . The following changes are made: magnetic
10 1. Plasma: an ionized gas
eld B B/c, vector potential A A/c, permittivity of free space
0
1/4, permeability
of free space
0
4/c
2
. Other conversion factors are include in the following Table 1.3. Some
other conversion factors are given in Table 1.4.
Table 1.3: Conversion factors between SI and gaussian units
quantity SI/gaussian gaussian/SI
length 10
2
m/cm 10
2
cm/m
mass 10
3
kg/g 10
3
g/kg
energy 10
7
J/erg 10
7
erg/J
power 10
7
W/erg s
1
10
7
erg s
1
/W
force 10
5
N/dyne 10
5
dyne/N
charge
1
3
10
9
statcoul/C 3 10
9
C/statcoul
electric eld 3 10
4
Vm
1
/statvolt cm
1
1
3
10
4
statvolt cm
1
/Vm
1
current
1
3
10
9
A/statamp 3 10
9
statamp/A
current density
1
3
10
5
Am
2
/statamp 3 10
5
statamp cm
2
/Am
2
magnetic induction 10
4
T/G 10
4
G/T
Table 1.4: Other conversion factors
quantity factor inverse factor
temperature 8.6 10
5
eV/K 1.16 10
4
K/eV
X-ray energy 4.1 10
15
eV/Hz 2.4 10
14
Hz/eV
angle 2.06 10
5
arcsec/rad 4.85 10
6
rad/arcsec
time 3.16 10
7
s/yr 3.17 10
8
yr/s
1.8 Exercise Set 1 11
1.8 Exercise Set 1
1.1 Calculate the plasma frequency, the electron cyclotron frequency for the following parameters:
(a) A laboratory fusion plasma: n
e
= 10
15
m
3
, B = 1 T, T
e
= 10 keV.
(b) The ionosphere: n
e
= 0.1 m
3
, B = 10
5
T, T
e
= 10
3
K.
(c) The solar corona: n
e
= 10
10
cm
3
, B = 1 G, T
e
= 10
6
K.
(d) The interplanetary medium: n
e
= 1 cm
3
, B = 3 10
6
G, T
e
= 10
6
K.
1.2 Calculate the Debye length and the Debye number for the plasmas in Exercise 1.1.
1.3 Calculate the Alfven speed for the plasmas in Exercise 1.1.
12 1. Plasma: an ionized gas
Chapter 2
Collective eects in plasmas
A characteristic feature of a plasma is its collective response. The charges in the plasma move
in response to electric and magnetic elds, creating induced charge and current densities in the
plasma. The plasma frequency characterizes the response of the plasma to temporal oscillations of
an electric eld, and the Debye length characterizes the spatial response to an electrostatic eld.
In this lecture these purely temporal and purely spatial responses are discussed separately.
2.1 Electrostatic oscillations in cold plasma
Suppose one applies an oscillating eld, E(t), across a plasma. The electric eld accelerates
electrons in one direction and ions in the opposite direction. Due to their smaller mass, the
electrons move more freely than the ions. For simplicity, let us ignore the motion of the ions,
assuming that they have eectively innite mass, or are tied in some crystalline structure, and
concentrate only on the electrons. Furthermore, suppose the plasma is conned to a slab with
surfaces perpendicular to E. The electric eld causes the electrons to be displaced from the ions,
by a distance d(t) say along the direction of E(t). This does not change the charge density inside
the plasma, but it induces a surface charge density. On one side of the slab there is an excess of
electrons, and hence a negative charge, and on the other side there is a deciency of electrons and
hence a positive charge, as illustrated in Figure 2.1. This gives a surface charge (t) = en
e
d(t)
on one side and (t) = +en
e
d(t) on the other side. A surface charge generates and electric
eld /2
0
of opposite sign on opposite sides of the surface. Inside the plasma, the electric elds
from the surface charge densities on either side add, giving an electric eld en
e
d(t)/
0
inside the
14 2. Collective eects in plasmas
d d
+
+
-
+
-
-
Figure 2.1: A schematic showing the electrons (dashed region) separated from the ions (solid
region) by a distance d, setting up a net charge density that is negative where there is an excess
of electrons, and positive where there is a deciency of electrons.
plasma. This induced electric eld adds to the applied eld, modifying it so that it becomes a
self-consistent eld.
Now suppose that there is no applied eld. The only electric eld is then due to free oscillations
of the electrons relative to the ions. Any displacement of the electrons from the ions sets up the
internal eld E(t) = en
e
d(t)/
0
. This electric eld accelerates the electrons. Newtons equation
of motion implies
m
e

d(t) = eE(t) = e
2
n
e
d(t)/
0
, (2.1)
where a dot denotes a derivative with respect to time. Equation (2.1) is the equation for a simple
harmonic oscillator with frequency equal to the plasma frequency,
p
= (e
2
n
e
/m
e

0
)
1/2
. To see
this, suppose that d(t) = d
0
cos(t) is oscillating at frequency with amplitude d
0
. On inserting
this into (2.1), one has

d(t) =
2
d(t), and (2.1) reduces to m
e

2
d(t) = e
2
n
e
d(t)/
0
, which
implies
2
=
2
p
. It follows that the electron plasma frequency is the natural frequency of free
oscillations of the electrons relative to (immobile) ions.
2.2 Fluid model for electron gas 15
2.2 Fluid model for electron gas
Another way of deriving the plasma frequency is to regard the electron gas as a uid, described
in terms of its density, pressure and uid velocity. The number density, n
e
(x, t) = n
e
+ n
e
(x, t),
and the pressure, P
e
(x, t) = P
e
+P
e
(x, t), vary in space and time about mean values, n
e
and P
e
,
respectively. The mean values of the uid velocity u
e
(x, t) and of the electric eld, E(x, t), are
assumed to be zero. We need two uid equations.
One uid equation is the continuity equation

t
n
e
(x, t) + div [n
e
(x, t) u
e
(x, t)] = 0, (2.2)
which expresses conservation of electrons. This may be seen by integrating (2.2) over a volume,
V , with a surface, S. The rst term in (2.2) integrates to give the rate of change of the number
of electrons in V . The second terms may be rewritten using Stokes theorem, so that it becomes
the rate of escape of electrons across S. Together these express conservation of electrons.
The other uid equation is the equation of uid motion
m
e
_

t
+u
e
(x, t)

x
_
u
e
(x, t) = grad P
e
(x, t) e[E
e
(x, t) +u
e
(x, t) B(x, t)], (2.3)
The left hand side is the rate of change of the momentum density of the electrons, with the second
term involving the so-called convective derivative. The right hand side represents the force per
unit volume acting on the electron uid. The rst terms is the pressure gradient, the second term
is the electric force, and the next term is the Lorentz force. One could also include a mechanical
force, such as gravity but we do not do so here.
Here we are interested in electrostatic eld, with B(x, t) = 0, and
div E(x, t) = en
e
(x, t)/
0
. (2.4)
A further simplifying assumption is that the uid velocity, the electric eld and the variations
in the electron density and pressure are small, so that only terms linear in small variations are
retained. Then (2.2) and (2.3) give

t
n
e
(x, t) + n
e
div u
e
(x, t) = 0, m
e

t
u
e
(x, t) = eE
e
(x, t), (2.5)
respectively, where the pressure force is also neglected for simplicity. On taking the time derivative
of the rst of equations (2.5), taking the divergence of the second of equations (2.5), and using
16 2. Collective eects in plasmas
(2.4) with n
e
(x, t) replaced by n
e
(x, t), one nds that n
e
(x, t) satises the oscillator equation
with a natural frequency
p
. Although the mathematical details are quite dierent, this derivation
is physically equivalent to that given above based on a slab model.
Purely temporal oscillations may be regarded as a limiting case of a wave motion in which
the wavelength is innite. Before considering the case where the wavelength is nite, we need to
consider the static limit, in which the electric eld is a function of space but is not changing as a
function of time. This corresponds to the limit of a wave with zero frequency or innite period.
2.3 Collective response to a static eld
A static distribution of charge in vacuo creates an electrostatic eld, which is a potential eld, and
so may be written as minus the gradient of a potential: E = grad . The eld and the charge
density are related by one of Maxwells equations
div E =
2
= /
0
. (2.6)
In particular, a point charge, q, at the origin, x = 0, creates a Coulomb eld,
(x) =
q
4
0
r
, E(x) =
q
4
0
x
r
3
. (2.7)
where r = [x[ is the radial distance from the origin. To prove that (2.7) is the solution of (2.6) for
a point charge involves introducing the Dirac -function. The charge distribution corresponding
to a point charge is described by a -function. For a charge q at x = x
0
, the charge density is
(x) = q
3
(x x
0
), with
3
(x x
0
) = (x x
0
)(y y
0
)(z z
0
).
A static distribution of charge in a plasma creates an electric eld which is dierent from that
in vacuo. Consider inserting a charge q > 0 at the origin. Through its Coulomb eld, the charge q
attracts negative charges and repels positive charges. The electrons in the plasma, moving around
like the molecules of air in this room, are attracted to the charge q, and this causes the number
density of electrons to be slightly higher near q. Similarly, the positive ions in the plasma are
repelled by q, and their number density is slightly lower near q. Thus, the presence of q induces
a charge density in the plasma. The total electric eld is the sum of the Coulomb eld and the
electric eld due to this induced charge density in the plasma.
Consider the eect of charges in a shell between r and r + dr. The number of particles in
this shell is the number density of particles times the volume, 4r
2
dr, of the shell. The force
2.4 Debye screening 17
(and its reaction) due to each of these particles decreases 1/r
2
, due to the inverse-square-
law form of the Coulomb interaction, but this is oset by the number of particles in the shell
increasing r
2
. Summing over the shells (integrating over dr) the result diverges. This leads to
the surprising conclusion that the eld associated with a test charge, q, in a plasma is modied
due to interactions with all the other charges in the plasma. This situation is quite dierent to the
case of an un-ionized gas. The (van der Waals) force between molecules falls o 1/r
6
, and only
the interactions with nearest neighbors are important, with the net eect of other particles falling
o rapidly, 1/r
4
. However, in a plasma there is no decrease in the net force, and one cannot
assume that the eect of nearest neighbors dominates. To overcome this diculty we introduce
the concept of a self-consistent eld.
It is somewhat easier to understand the concept of a self-consistent eld in a related context:
a cluster of stars. The gravitation attraction between stars is an inverse-square-law force, and the
same problem arises: all the stars in the cluster aect the motion of any one star. We model the
cluster by a smoothed gravitational potential and a smoothed mass density, and relate these by
Poissons equation. Once we nd this self-consistent eld, we consider the motion of an individual
star in the smoothed potential. In the electrostatic case, there are charges of opposite sign, and
the self-consistent eld leads to Debye screening.
2.4 Debye screening
Debye screening may be treated in the following approximate way. The Coulomb eld (2.7) is the
solution of
div E =
2
= /
0
, = q
3
(x), (2.8)
corresponding to a point charge at the origin x = 0, with r = [x[ the radial distance from the
origin. The eect of this eld on surrounding charges is to attract the charges of opposite sign to
q, and repel charges of the same sign as q. Consider the thermal electrons in the plasma. Their
number density is modied such that it becomes by
n
e
(r) = n
e
exp[e(r)/T
e
], (2.9)
18 2. Collective eects in plasmas
where n
e
is the mean electron number density, and e(r) is the potential energy of an electron
in the potential (r). The perturbation introduces an additional charge density in the plasma
= e[n
e
(r) n
e
] e
2
n
e
(r)/T
e
, (2.10)
where [e(r)[ T
e
is assumed. This additional charge density needs to be included in (2.8),
which becomes

2
(r) = (r)/
2
D
+ (q/
0
)
3
(x), (2.11)
with e
2
n
e
/
0
T
e
= 1/
2
D
. The solution of (2.11) is
(r) =
q
4
0
r
e
r/
D
. (2.12)
Thus, the Coulomb eld is shielded or screened out at distances r
D
.
2.5 Plasma parameter
A qualitatively important quantity is the plasma parameter, or the Debye number. It may be
dened as the number of particles within a Debye sphere:
N
D
=
4
3
n
e

3
D
. (2.13)
The natural logarithm of the Debye number, usually written ln is called the Coulomb logarithm.
The collective eects that characterize a plasma require that the Debye number be much
greater than unity. To see this, note that screening requires small perturbations in the motions of
a large number of particles due to the Coulomb eld. For the Coulomb eld to be screened out
for r
D
required a large number of particles at r
D
. For N
D
1, there are eectively no
particles at r
D
, and the concept of screening by non-existent particles is meaningless. The
concept of the collective response of a plasma requires N
D
1
2.6 Exercise Set 2 19
2.6 Exercise Set 2
2.1 Calculate the electric eld corresponding the the screened potential (2.12). Show that your
result reduces to a Coulomb eld for
D
.
2.2 Estimate N
D
for the following plasmas: (a) an arc discharge with n
e
= 10
25
m
3
, T
e
= 1 eV,
(b) the solar corona with n
e
= 10
10
cm
3
, T
e
= 10
6
K, (c) an interstellar cloud, with n
e
= 1 cm
3
,
T
e
= 100 K.
2.3 Show that the Coulomb eld, = q/4
0
r is a solution of
2
= q
3
(x) by integrating

2
= q
3
(x) over a sphere of radius r centered on the charge, and noting that the right hand
side gives q. Note that in spherical polar coordinates, r, , , one has

2
=
1
r
2

r
r
2

r
+
1
r
2

2
(cos )
2
+
1
r
2
sin
2

2
.
20 2. Collective eects in plasmas
Chapter 3
Wave dispersion in plasmas
Plasmas support a rich variety of dierent types of waves, called wave modes. There is no system-
atic way of naming wave modes: some are historical (Langmuir waves, Alfven waves), some are
descriptive of the waves (transverse waves, cyclotron waves) and some are description of the theory
used to describe them (magnetoionic waves, MHD waves). Some wave modes can be described
using relatively simple models, but this is the case only for the simplest systems. In this lecture
and the next lecture, a general description of a wave mode in a medium is introduced, and some
properties of waves in plasmas are discussed.
3.1 Sound waves
A simple example of a waves is a sound wave in an un-ionized compressible gas. Such waves exist
only at very low frequencies in a plasma, specically at frequencies well below the collision frequen-
cies between particles. Nevertheless, the example of sound waves serves as a useful introduction
to waves in plasmas.
It is important to distinguish between a plane wave, which is a mathematical construction,
and a physical wave in a medium. A plane wave is dened to vary in time in space harmonically,
proportional to exp[i(t k x)], where is the frequency and k is the wave vector. The plane
wave has a period, T, and a wavelength, , with = 2/T and k = 2/. The wave propagates
in the wave-normal direction, described by the unit vector say, with k = k. A plane wave
is an idealization, and a real disturbance may be regarded as a superposition of plane waves.
Technically, a Fourier transform in time and space corresponds to an expansion in plane waves.
22 3. Wave dispersion in plasmas
We describe physical waves in a medium in terms of a solution of the appropriate wave equation.
Given a model for a specic medium we can derived the wave equation from the equations for the
model.
Consider the example of a sound wave. The model in this case is a compressible uid, described
by the equations of hydrodynamics. The wave equation for sound waves is derived as follows. A
sound wave corresponds to a uctuation in the gas, and this may be described by a uctuation,
u, in the velocity of the gas. There are also uctuations in the mass density, =
0
+
1
and in
the pressure, P = P
0
+ P
1
, where subscript 0 implies that non-uctuating part, and subscript 1
describes the uctuating part. The adiabatic law for a perfect gas is
P

= constant, (3.1)
where the adiabatic index is = 5/3 for a monatomic gas. This implies that the uctuations in
the density and pressure are related by
P
1
= c
2
s

1
, c
2
s
= P
0
/
0
, (3.2)
where c
s
is the (adiabatic) sound speed. (The derivation of (3.2) involves linearizing (3.1): one
has (
0
+
1
)

0
(1
1
/
0
) to rst order.) The hydrodynamic equations are the continuity
equation for mass
/t + div (u) = 0, (3.3)
where u is the uid velocity, and the equation of uid motion
du/dt = grad P. (3.4)
One assumes that the amplitude of the uctuations,
1
, P
1
, u, are small, so that products of them
can be neglected. Then taking the time derivative of (3.3) and the divergence of (3.4), one nds
that
1
, P
1
, u all satisfy the same wave equation. For P
1
this is
[
2
/t
2
c
2
s

2
]P
1
= 0. (3.5)
A solution of (3.5) corresponds to a wave propagating at the speed c
s
.
3.2 Expansion in plane waves 23
3.2 Expansion in plane waves
We look for a solution of (3.5) that varies periodically, as cos(t k x) or sin(t k x). It is
convenient to introduce complex notation by writing
cos x =
1
2
(e
ix
+e
ix
), sin x =
1
2i
(e
ix
e
ix
),
and to choose to look for a solution that varies as exp[i(t k x)]. Thus in (3.5) we look for a
solution in which the variation of P
1
with time and position is exp[i(t k x)]. In a more
formal treatment, this step is replaced by Fourier transforming in space and time.
On expanding the uctuations in plane waves, the derivatives in (3.5) operate only on exp[i(t
k x)]. One has

t
exp[i(t k x)] = i exp[i(t k x)],
When the time derivative and the derivatives with respect to position operate only on the expo-
nential function, they are replaced according to

t
i, grad = ik, div = ik, curl = ik, (3.6)
with
2
= div grad [k[
2
. The dierential operator in (3.5) is replaced according to

2
t
2
c
2
s

2
+k
2
c
2
s
.
It follows that a plane-wave solution of (3.5) exists only for

2
k
2
c
2
s
= 0. (3.7)
The relation = kc
s
is referred to as the dispersion relation for sound waves.
The simple model enables one to infer other properties of sound waves. One can show that the
uid velocity is parallel to the wave normal direction, and that there is equipartition between the
average kinetic energy density,
1
2

0
[u[
2
, and the average potential energy,
1
P
1
/2
0
, in the wave.
Other wave modes
The uid model can be generalized in various ways, and each generalization leads to modication
of the properties of the wave modes, and to the appearance of new wave modes. Suppose one
considers the Earths atmosphere, and takes the decrease in the density with height, z, in account.
24 3. Wave dispersion in plasmas
In an isothermal model, the decrease is of the form exp(z/H), where the scale height, H = c
2
s
/g
depends on the acceleration due to gravity, g. This aects the properties of sound waves, which do
not exist at frequencies <
1
,
2
1
= c
2
s
/4H
2
. No waves can propagate in the range
2
< <
1
,
with
2
2
= ( 1)g
2
/c
2
s
. Below
2
, called the Vaisala-Brunt frequency, a dierent kind of wave,
called an internal gravity wave, can propagate. At
2
the dispersion relation for internal
gravity waves can be approximated by = (kg)
1/2
. Another generalization of the uid theory is
to include a magnetic eld and assume that the medium is electrically conducting. This leads to
MHD waves. There are three MHD wave modes. One is a modied sound wave, and the other
two are Alfven waves, associated with the tension in the magnetic eld lines, and magnetoacoustic
waves, associated with the magnetic pressure. The dispersion relations for the MHD waves are
derived in a later lecture.
Dierent types of waves are referred to as dierent wave modes. Thus one refers to the sound
mode, the Alfven mode, and so on. A characteristic property for a wave mode is its dispersion
relation, which can be written as an algebraic relation between and k. It is useful to introduce
the concept of an arbitrary wave mode, which we refer to as the mode M. The dispersion relation
for the mode M can always be written in a form =
M
(k), that is, the frequency as a function
of the wave vector. For sound waves, this notation corresponds to a dispersion relation =
s
(k),
where the label s is for sound and with
s
(k) = kc
s
. For electromagnetic waves in vacuo, it is
convenient to use the label T, for transverse, with the dispersion relation =
T
(k) corresponding
to
T
(k) = kc. For transverse waves in a medium, the dispersion relation is often written in terms
of the refractive index, n = kc/, with n = 1 for transverse waves in vacuo. Transverse waves in
a plasma have a refractive index n = (1
2
p
/
2
)
1/2
, which corresponds to
T
(k) = (
2
p
+k
2
c
2
)
1/2
,
where
p
is the plasma frequency.
3.3 Phase and group velocities
There are two important velocities associated with the propagation of a wave: the phase velocity
and the group velocity.
3.3 Phase and group velocities 25
Phase velocity: particle-wave resonance
The phase velocity is the velocity at which surfaces of constant phase move. The surfaces of
constant phase correspond to t k x = constant. The phase velocity is dened to be along the
direction of k, that is, along the wave normal direction. The surfaces of constant phase move in
this direction at the phase speed v

= /k. Thus the phase velocity for a wave in the mode M is

M
(k)/k along k. There is no physical restriction on the phase velocity, which can be greater or
less than the speed of light. Waves with v

> c (v

< c) are said to be superluminal (subluminal).


The phase velocity plays an important role in plasma physics through wave-particle resonance.
A particle whose velocity v is equal to the phase velocity of a wave satises the resonance condition
k v = 0. (3.8)
One can understand the importance of the resonance condition by considering what an observer
moving with the particle sees. This observer sees a particle at rest and a wave that varies periodi-
cally in space but is stationary in time. Plasma waves include an electric eld, and the particle is
systematically accelerated by this electric eld. Returning to the frame in which the particle has
velocity v, one nds that particles near resonance tend to be dragged into resonance by this eect.
Particles with velocity slightly less than that of the wave gain energy at the expense of the wave,
and particles with velocity slightly greater than that of the wave give up energy to the wave, as
they are dragged into resonance with the wave. In a thermal distribution, the number of particles
with a given velocity is determined by a Maxwellian distribution exp(mv
2
/2T), where T is
the temperature (in energy units). There are then more particles with speed slightly less than a
given speed than particles with speed slightly greater than this speed, and the particles gain a net
energy from the wave. This causes the wave to damp, through what is called Landau damping.
A non-Maxwellian distribution may have a distribution function which is an increasing function
of velocity over some range, and in this case waves with phase velocity in the range gain energy
from the wave. This leads to growth of waves in what is called a plasma instability.
Group velocity
The energy in waves propagates at the group velocity. The group velocity for the mode M is
v
gM
=
M
(k)/k. (3.9)
26 3. Wave dispersion in plasmas
For transverse waves in a cold plasma, one has
v
gT
=
_

2
p
+k
2
c
2
/k = kc
2
/
_

2
p
+k
2
c
2
= c(1
2
p
/
2
)
1/2
, (3.10)
where = k/k is the wave-normal direction.
In an anisotropic medium, the group velocity and the phase velocity are in dierent directions
in general. In terms of cartesian components, with k = (k
x
, k
y
.k
z
), the cartesian components of
the group velocity are (/k
x
, /k
y
, /k
z
)
M
(k).
Ray propagation
The path along which the wave energy propagates corresponds to the ray path. In a medium
whose properties change slowly with space and time, the dispersion relation changes slowly with
space and time. Writing
M
(k; t, x) to include this slow change, the path of a ray is determined
by the ray equations
dx
dt
=

k

M
(k; t, x),
dk
dt
=

x

M
(k; t, x),
d
dt
=

t

M
(k; t, x). (3.11)
The rst of (3.11) implies that the velocity along the ray path is the group velocity. Suppose that
the properties of the medium are varying along the z direction. Then the second of (3.11) implies
that k
z
is changing along the ray path, and that the components, k
x
, k
y
, perpendicular to this
direction are constant, which is Snells law.
3.4 Exercise Set 3 27
3.4 Exercise Set 3
3.1 Show that the equations (3.1), (3.3) and (3.4), with d/dt = /t +u grad, imply

t
_
[u[
2
2
+
P
1
_
+ div
_
[u[
2
2
u +
P
1
u
_
= 0, (3.12)
which is interpreted as the equation for energy continuity for the uid. The rst two terms are the
kinetic energy density and the thermal energy density, and the second two terms are the kinetic
energy ux, and the thermal energy ux, with the latter equal to the enthalpy times u.
3.2 Show that in a sound wave, there is equipartition between the kinetic energy density, W
K
=
1
2

0
[u[
2
, and the potential energy density, W
P
=
1
2

2
1
c
2
s
/
0
, associated with the pressure uctua-
tions.
3.3 Derive the sound speed in the case where the medium is assumed isothermal rather than
adiabatic.
3.4 The dispersion relation for Langmuir waves may be approximated by
2
L
(k) =
2
p
+ 3k
2
V
2
e
,
where V
e
is the thermal speed of electrons. Show the the phase speed, v

, and the group speed,


v
g
, for waves with this dispersion relation satisfy v

v
g
= 3V
2
e
.
3.5 Consider a model for Langmuir waves in which the electrons are treated as a compressible gas
with pressure P = n
e
m
e
V
2
e
satisfying the adiabatic law P n

e
where is the adiabatic index.
(a) Assuming an equation of uid motion of the form m
e
n
e
du/dt = en
e
E gradP, show that
the implied dispersion relation for Langmuir waves is

2
=
2
p
+k
2
V
2
e
. (3.13)
Hint: Use a rst order perturbation treatment with the zeroth order corresponding to P = 0.
Assume k u.
(b) This model reproduces the correct form (3.13) for = 3. Can this value of be justied, or
is the model inadequate to describe Langmuir waves?
28 3. Wave dispersion in plasmas
Chapter 4
Waves in isotropic plasmas
Plasmas contain charged particles whose motion is aected by electric and magnetic elds. The in-
duced charge and current densities associated with the collective eect of these perturbed motions
modies the electric and magnetic elds in the plasma, leading to the concept of a self-consistent
eld. Waves in plasma involve oscillations in these elds, and the natural wave modes of the plas-
mas may be regarded as free oscillation of the self-consistent eld. The wave equation is derived
by starting with Maxwells equations, and including the induced charge and current densities in
the plasma as responses to external or extraneous charges and currents that act as source terms.
4.1 Wave equation for a plasma
Maxwells equation are
curl E = B/t, div B = 0,
curl B =
0
J + (1/c
2
)E/t, div E = /
0
. (4.1)
All quantities in (4.1) are functions of t, x.
In applying Maxwells equation to a plasma, it is convenient to use the Fourier transforms of
(4.1). The Fourier transform is eectively an expansion in plane waves. Suppose one expands a
function, F, in plane waves, and writes it as an integral over terms of the form

F exp[i(tkx)],
where

F depends on , k. The dierential operators in (4.1) operate only on the exponential
function. The eect of the operators may be summarized by (3.6). As a result of expanding
30 4. Waves in isotropic plasmas
Maxwells equations (4.1) in plane waves and using (3.6), the Fourier components satisfy
k

E =

B, k

B = 0, k

B = i
0

E/c
2
, k

E = i /
0
, (4.2)
respectively. One may regard the rst of the equations (4.2) as determining

B in terms of

E. The
second equation is implied by the rst. The fourth equation, combined with the third, may be
interpreted as determining the charge density in terns of the current density,
= k

J. (4.3)
The third of equations (4.2) then becomes
k [k

E] + (
2
/c
2
)

E = i
0

J, (4.4)
which is one form of the wave equation.
Note what we have done: we have eectively reduced the four equations (4.1) to a single
equation (4.4). We achieve this by rst noting that div B = 0 is redundant for elds that are
varying in time, because it is implies by the rst of (4.1), so that our four equations are reduced
to three by expanding in plane wave, which ignores the static elds. Two of the remaining three
equations are regarded as subsidiary equations: (4.2) dening

B in terms of

E, and (4.3) dening
in terms of

J. The wave equation (4.4) relates the remaining quantities,

E and

J.
4.2 Wave equation
The next step is a particularly important one, in that it is the essential step in introducing the
self-consistent eld. First, let us separate the current into an induced (ind) part and an extraneous
(ext) part, by writing

J =

J
ind
+

J
ext
, (4.5)
We regard

J
ext
as a source term and leave it on the right hand side; it is set to zero when considering
the wave modes of the plasmas. The important assumption is that the response of the medium
may be described in terms of a linear relation between the component of

J
ind
and the components
of

E. One transfers

J
ind
to the left hand side of (4.4), so that all the terms proportional to the
components of

E are on the same side of the equation. With this assumption, (4.4) becomes three
4.3 Longitudinal response 31
simultaneous equations for the three components of

E. Let the components be labeled 1, 2, 3.
(Alternatively, we could write them as x, y, z.) The three equations can be written in matrix form
_
_
_
_
_
_
_

11

12

13

21

22

23

31

32

33
_
_
_
_
_
_
_
_
_
_
_
_
_
_

E
1

E
2

E
3
_
_
_
_
_
_
_
= 0, (4.6)
where we assume

J
ext
, so that there is no source term. The matrix components
ij
, with i, j =
1, 2, 3, are functions of and the components k
1
, k
2
, k
3
of k. An alternative way of writing (4.6)
is

ij

E
j
= 0, (4.7)
where the sum over the repeated index, j = 1, 2, 3, is implicit (Einsteins summation convention).
Equation (4.6) is a matrix equation for the three components of the electric eld. The condition
for a matrix equation to have a solution is that the determinant of the matrix of coecients vanish.
Thus the condition for a solution of (4.6) to exist is
(, k) = det [
ij
] = 0. (4.8)
Equation (4.8) is called the dispersion equation. A particular solution of (4.8), for as a function
of k say, is interpreted as the dispersion relation for a wave mode.
4.3 Longitudinal response
First let us concentrate on the longitudinal response. Longitudinal means the component of a
vector along k. The longitudinal component of the current is related to the charge density, and
the longitudinal part of the electric vector is related to the electrostatic potential,

. Hence, when
discussing the longitudinal response, we are free to describe it in terms of the charge density induce
by the electrostatic potential.
It is convenient to change notation, omitting the tilde, and writing the uctuating quantities
as functions of and k. One has
(, k) = k J(, k)/, (, k) = ik E(, k)/[k[
2
. (4.9)
After expanding in plane waves, Poissons equation becomes
[k[
2
(, k) = (, k)/
0
. (4.10)
32 4. Waves in isotropic plasmas
In the absence of a plasma, (4.10) gives the potential generated by a given charge distribution.
In this case, the only charge density is the extraneous one,
ext
(, k), which is the source of the
electric eld.
In the presence of a plasma, the extraneous charge distribution causes an induced charge
distribution, by attracting charges of one sign and repelling charges of the opposite sign. Let
the charge density associated with this induced response be
ind
(, k). The induced response is
assumed to be a linear function of the eld causing it, and here this implies
ind
(, k) (, k)

ext
(, k). The total charge density,
tot
(, k), in the plasma is the sum of the two contributions.

ext
(, k) and
ind
(, k). This implies that
tot
(, k) is proportional to the source term,
ext
(, k).
The constant of proportionality is called the longitudinal response function:

tot
(, k) = K
L
(, k)
ext
(, k). (4.11)
The response function, K
L
(, k), includes a unit term, corresponding the the extraneous charge
density on the right hand side of (4.11), and another term that describes the induced response of
the plasma.
Now let us reconsider (4.10). If we interpret (4.10) as an equation in a vacuum, then it gives
the potential due to the extraneous charge, and this corresponds to K
L
(, k) 1, describing the
response of the vacuum. If we interpret (4.10) as an equation for a plasma, then it includes the
induced current, and can be rewritten as
[k[
2
K
L
(, k)(, k) =
ext
(, k)/
0
. (4.12)
When considering the natural modes of the plasma, one ignores the source term. This implies
ignoring the right hand side in (4.12). The resulting wave equation is
K
L
(, k)(, k) = 0, (4.13)
which is the longitudinal counterpart of the wave equation (4.7). The condition for a solution of
(4.13) to exist is
K
L
(, k) = 0, (4.14)
which is the longitudinal counterpart of the dispersion equation (4.8).
A detailed calculation of K
L
(, k) involves kinetic theory. There are two simple limiting cases
K
L
(, k)
_
_
_
1 + 1/[k[
2

2
D
for 0,
1
2
p
/
2
for [k[ 0.
(4.15)
4.4 Cold electron gas 33
The rst of these is derived from a model for Debye screening, and the second follows from a
model for a cold electron gas.
4.4 Cold electron gas
Now let us general to include both the longitudinal and transverse parts, but ignore the thermal
motions (so that there is no pressure term in the equation of motion for the electrons). This
leads to the cold plasma model, which is the simplest useful model for a plasma. The response
of a plasma tis described by the relation between the induced current

J and the electric eld

E,
where we now revert to the notation used in expanding in plane waves. We wish to calculate the
elements in the matrix
ij
in (4.6) for a cold electron gas.
The current density in a cold electron gas is J = en
e
u. The only force is due to the
electromagnetic eld. The equation of motion for the uid is the same a Newtons equation
m
e
du
dt
= e[E +u B]. (4.16)
The Lorentz force, u B, is the product of two uctuating quantities, and so is neglected. After
expansing in plane waves, (4.16) gives
im
e
u = e

E. (4.17)
The current density becomes

J = i
e
2
n
e
m
e

E = i
0

2
p

E. (4.18)
On inserting (4.18) into (4.4) the wave equation becomes, after minor rearrangement,
(
2

2
p
k
2
c
2
)

E +c
2
kk

E = 0. (4.19)
It is convenient to divide by
2
, to introduce the refractive index by writing n
2
= k
2
c
2
/
2
, and to
write k = k. Then (4.19) can be written in the matrix form

E = 0, = n
2
[ 1] +1 K(), K() = 1

2
p

2
, (4.20)
where 1 is the unit matrix and K() is dielectric constant for a cold plasma.
34 4. Waves in isotropic plasmas
We are free to choose the coordinate axes such that k is along the 3-axis. Then (4.19) can be
written in the form (4.6), with the matrix being diagonal. One can write the result in the form
_
_
_
_
_
_
_

T
0 0
0
T
0
0 0
L
_
_
_
_
_
_
_
_
_
_
_
_
_
_

E
1

E
2

E
3
_
_
_
_
_
_
_
= 0, (4.21)

L
= K
L
,
T
= K
T

k
2
c
2

2
, K
L
= K
T
= 1

2
p

2
. (4.22)
The result (4.22) for K
L
is the second of the two simple cases written down in (4.15).
The dispersion equation (4.8) becomes
(, k) =
L
[
T
]
2
= 0. (4.23)
The solution
L
= 0 corresponds to the component

E
3
,= 0 with

E
1
=

E
2
= 0, and the solution

T
= 0 corresponds to the component

E
3
= 0 with

E
1
,

E
2
,= 0. These are referred to as longitu-
dinal waves, with

E parallel to k, and transverse waves, with

E orthogonal to k, respectively.
The longitudinal waves are the electron plasma oscillations, at =
p
, rst identied by Lang-
muir. When thermal motions are included, the dispersion relation for Langmuir waves becomes
=
L
(k), with
2
L
(k)
2
p
+ 3k
2
V
2
e
.
The dispersion relation for transverse waves in a cold electron gas is
T
= 0. This may be
written either in term of the refractive index, n = kc/, or in the form =
T
(k), with
n
2
= 1

2
p

2
,
2
T
(k) =
2
p
+k
2
c
2
. (4.24)
Transverse waves do not exist for <
p
in a cold plasma.
4.5 Energetics in waves
The total energy in a wave may be separated into electric energy, magnetic energy and kinetic
energy associated with the perturbed motions of the particles. The ratio of magnetic to electric
energy follows from the rst of equations (4.2):
W
M
: W
E
=
[

B
2
[
2
0
:

0
[

E
2
[
2
= k
2
c
2
:
2
. (4.25)
4.5 Energetics in waves 35
In general it is not possible to calculate the kinetic energy in the waves directly, but the cold
plasma model is an exception. In this case, the ratio of the kinetic energy to the electric energy
follows from (4.17)
W
K
: W
E
=
n
e
m
e
[ u
2
[
2
:

0
[

E
2
[
2
=
2
p
:
2
. (4.26)
Hence, for a transverse wave in a cold plasma the total energy is made up from these contributions
in the ratio
W
E
W
T
:
W
M
W
T
:
W
K
W
T
=
1
2
:
1
2
_
1

2
p

2
_
:
1
2

2
p

2
, (4.27)
where W
T
= W
E
+ W
M
+ W
K
is the total energy in the waves. For Langmuir waves there is no
magnetic energy, and there is approximate equipartition between electric and kinetic energy.
The propagation of a wave implies propagation of energy at the group velocity. The phase
velocity is /k, or
M
(k)/k for the mode M. The group velocity, (3.9), viz. v
gM
=
M
(k)/k,
for transverse waves in a cold plasma is
v
gT
=
_

2
p
+k
2
c
2
/k = kc
2
/
_

2
p
+k
2
c
2
= c(1
2
p
/
2
)
1/2
. (4.28)
In a cold plasma, the velocity of energy propagation is also given by the Poynting ux, EB/
0
,
divided by the total energy density in the waves. Using the rst of equations (4.2), the Poynting
ux becomes

E

B

0
=
kc
2


0
[

E[
2
, (4.29)
and the total energy is W
E
+W
M
+W
K
= 2W
E
=
0
[

E[
2
for transverse waves.
Langmuir waves propagate very slowly,
v
gL
=
_

2
p
+ 3k
2
V
2
e
/k 3kV
2
e
/
p
. (4.30)
Their group velocity reduces to zero in the cold plasma limit.
36 4. Waves in isotropic plasmas
4.6 Exercise Set 4
4.1 The two states of polarization for transverse waves are degenerate in an isotropic dielectric,
an isotropic plasma or the vacuum. The solution of (4.7) allows E to be in any direction in the
plane orthogonal to k. With k along the z-axis, E = (E
x
, E
y
, 0) can have any (complex) values
E
x
, E
y
. If we normalize to unity, and write the normalized solution as the polarization vector e,
with e e

= 1, then we have [e
x
[
2
+ [e
y
[
2
= 1. The polarization vector is dened only to within
an arbitrary phase factor. We may choose this phase factor such that the polarization vector is
e =
1
(1 +T
2
)
1/2
(cos iT sin , sin +iT cos , 0), (4.31)
where is an arbitrary angle, and T is an arbitrary real number. Two polarization vectors, e
(1)
,
e
(2)
say, are orthogonal if they satisfy e
(1)
e
(2)
= 0.
(a) Show that two polarization vectors, e
+
, e

are orthogonal if they have the same and T = T

,
with T
+
T

= 1.
(b) Determine the values of and T for linear polarization along (i) the x axis, and (ii) the y axis.
(c) Determine the values of and T for right- and left-hand circular polarization:
E
r,l
=
1

2
(1, i, 0). (4.32)
(d) Show that an arbitrary polarization is elliptical, with T the axial ratio of the ellipse, and with
the sign of T determining whether the electric vector rotates in a right-hand or left-hand sense
around the ellipse.
4.2 Certain isotropic dielectric, such as a solution of dextrose are optically active in the sense
that when linearly polarized light propagates through the medium, the plane of linear polarization
rotates at a characteristic rate along the ray path. The most general isotropic medium can be
described by three dielectric constants, K
L
, K
T
and K
R
, with K
R
= 0 in media that are not
optically active. When K
R
is nonzero, (4.21) with (4.22) is replaced by
_
_
_
_
_
_
_
K
T
k
2
c
2
/
2
iK
R
0
iK
R
K
T
k
2
c
2
/
2
0
0 0 K
L
_
_
_
_
_
_
_
_
_
_
_
_
_
_

E
1

E
2

E
3
_
_
_
_
_
_
_
= 0. (4.33)
(a) Write down the dispersion equation by setting the determinant of the square matrix in (4.33)
to zero.
4.6 Exercise Set 4 37
(b) Show that there are two modes, and nd the values of n
2
= k
2
c
2
/
2
for these two modes.
(c) Show that the polarization vectors of the modes correspond to circular polarizations.
4.3 The most general form of polarization for transverse waves in an isotropic medium (or the
vacuum) consists of an unpolarized component, and a polarized component, with the latter cor-
responding to an elliptical polarization in general. Let r be the degree of polarization, with r = 1
for completely polarized radiation, and in Exercise 4.1, and with r = 0 for unpolarized radia-
tion. Unpolarized radiation cannot be described by a polarization vector. One can describe it
by a polarization tensor, p
ij
. The polarization tensor is hermitian, p
ij
= p

ji
, and for completely
polarized radiation it reduces to the outer product of the polarization vector and its complex
conjugate: p
ij
e
i
e

j
. We are free to choose k along the z axis, and then p
ij
has components
that are zero for i or j equal to z. It is convenient to write it as a 2 2 matrix. We are free to
impose a normalization condition, and we require that the trace of the matrix be equal to unity:
p
xx
+p
yy
= 1. The most general form for an hermitian 2 2 matrix with this normalization is
p
ij
= (1 r)
ij
+r[p
Q
(
Q
)
ij
+ p
U
(
U
)
ij
+p
V
(
V
)
ij
], (4.34)
with p
2
Q
+p
2
U
+p
2
V
= 1, and with

ij
=
_
_
_
1 0
0 1
_
_
_, (
Q
)
ij
=
_
_
_
1 0
0 1
_
_
_, (
U
)
ij
=
_
_
_
0 1
1 0
_
_
_, (
V
)
ij
=
_
_
_
0 i
i 0
_
_
_, (4.35)
which are the unit matrix and the three Pauli matrices.
Show that
(a) r = 1, p
Q
= 1 correspond to linear polarization along the x, y axis, respectively;
(b) r = 1, p
V
= 1 correspond to right and left hand circular polarization, respectively;
(c) r = 1, p
U
= 0 corresponds to an elliptical polarization, with = 0 in (4.31), and express p
Q
,
p
V
in terms of the axial ratio, T.
38 4. Waves in isotropic plasmas
Chapter 5
Orbit theory
So far we have assumed that there is no background magnetic eld: the plasma is assumed
unmagnetized. We now include the magnetic eld. We discuss three topic that involve the
magnetic eld: orbit theory, cold plasma theory and magnetohydrodynamics.
Orbit theory is concerned with the motion of charged particles in a magnetic eld. Motion in
a uniform B is a spiraling along the magnetic eld lines. Drifts across the magnetic eld occur
in the presence of an electric eld, a mechanical force, and gradients in B. Conserved quantities,
known as adiabatic invariants, are helpful in understanding the motion.
5.1 Motion of a charged particle
Consider a particle with charge q and mass m moving in a magnetic eld B and an electric eld
E. Newtons equation of motion is
dp
dt
= q[E +v B], (5.1)
where v is the velocity of the particle and where
p = mv, = mc
2
, = (1 v
2
/c
2
)
1/2
(5.2)
are the momentum, energy and Lorentz factor of the particle, respectively.
In a uniform magnetic eld with no electric eld, E = 0, the following quantities are constants
of the motion: the energy = mc
2
, and the components p

= mv

and p

= mv

of the
momentum perpendicular and parallel to the magnetic eld, respectively. The motion of the
40 5. Orbit theory
particle may be decomposed into a motion at constant velocity along the eld lines plus a circular
motion perpendicular to the eld lines. Thus the particle exhibits a spiraling motion with the
pitch of the spiral dening the pitch angle :
v

= v sin , v

= v cos . (5.3)
The frequency of the circular motion is called the gyrofrequency, , and the radius is called the
radius of gyration (or sometimes the Larmor radius), R:
=

0

,
0
=
[q[B
m
, R =
v

=
p

[q[B
. (5.4)
The sense of gyration, which is the handedness of the circular motion in a screw sense relative to
B, depends on the sign of the charge
= q/[q[. (5.5)
Positively charged particles ( = +1) gyrate in a left hand screw sense relative to B, and negatively
charged particles ( = 1) gyrate in a right hand screw sense relative to B.
The orbit of the particle is described by an equation that gives the position x of the particle
as a function of time, and can be written as x = X(t). Solving (5.1) in this case gives
X(t) = x
0
+ (Rsin(
0
+ t), Rcos(
0
+ t), v

t), (5.6)
where
0
and x
0
are determined by the position of the particle at t = 0, and where the z axis is
chosen along the direction of B. The instantaneous velocity of the particle is given by
v(t) =

X(t) = (v

cos(
0
+ t), v

sin(
0
+ t), v

). (5.7)
The sense of gyration is such that the magnetic eld produced by the spiraling charge opposes the
externally applied eld. Plasmas are diamagnetic the motions of the individual particles always
tend to reduce the applied magnetic eld.
5.2 Electric drift
Now consider the eect of inclusion of a uniform, nonzero electric eld, E in (5.1). In most
applications it is assumed that there is no parallel component, i.e. E B = 0. The reason is that
plasmas are highly electrically conducting, and charges can ow freely along magnetic eld lines
5.2 Electric drift 41
to short out any parallel component of electric eld. However, particles do not ow freely across
eld lines and so an electric eld perpendicular to the magnetic eld is not shorted out and can
persist.
So far we choose the z axis to be along the direction of B, We are free to rotate our coordinate
system about this axis, and to choose this rotation such that E = (E, 0, 0). For nonrelativistic
particles the dierent components of equation (5.1) are then
dv
x
dt
=
q
m
E + v
y
,
dv
y
dt
= v
x
,
dv
z
dt
= 0. (5.8)
The equation for the z-component is trivial. Dierentiating the other two, using the fact that E
is constant, we obtain
d
2
v
x
dt
2
=
dv
y
dt
=
2
v
x
,
d
2
v
y
dt
2
=
dv
x
dt
=
2
_
v
y
+
E
B
_
. (5.9)
The equation for v
y
can then be rewritten as
d
2
dt
2
_
v
y
+
E
B
_
=
2
_
v
y
+
E
B
_
, (5.10)
and then if we make the replacement
v

y
= v
y
+E/B, (5.11)
equation (5.10) reduces to the same form as the equation for v
x
in (5.9). These are the equations
for a simple harmonic oscillator solved for the case E = 0. By analogy with equations (5.6) and
(5.7) we have the following orbit equations:
X(t) = x
0
+ (Rsin(
0
+ t), Rcos(
0
+ t) Et/B, v

t), (5.12)
v(t) = (v

cos(
0
+ t), v

sin(
0
+ t) E/B, v

), (5.13)
for the orbit and the instantaneous velocity, respectively.
An alternative way of understanding the eect of a perpendicular electric eld is to note that
the eld may be removed by making a Lorentz transformation. The quantities EBand B
2
E
2
/c
2
are Lorentz invariants, and provided E < B one may transform to a (primed) frame with E

= 0
and B

= (B
2
E
2
/c
2
)
1/2
. The velocity of the transformation is in the direction perpendicular
to both E and B, and is of magnitude v
E
= E/B. The motion of the particles in the primed
frame is a spiral around the magnetic eld B

. The fact that the primed frame drifts relative


42 5. Orbit theory
to the unprimed frame implies that the particle motion in the unprimed frame is a simple spiral
around a guiding center (or gyrocenter) which is drifting perpendicular to both the electric and
the magnetic elds. Note that in a uniform electric eld all particles drift with the same drift
velocity, v
E
.
5.3 Drift motions
A systematic treatment of drift motions involves assuming that the motion perpendicular to the
magnetic eld lines is circular motion about a center of gyration that is drifting. Eects that
cause drift include an electric eld, an external force, a gradient in the magnetic eld strength
and curvature of the eld lines. The following drifts are commonly identied:
electric drift : v
E
=
EB
B
2
, (5.14)
gravitational drift : v
g
=
mg B
qB
2
, (5.15)
gradient drift : v
B
=
p

2qB
BgradB
B
2
, (5.16)
inertial drift : v
i
=
B(dp/dt)
0
qB
2
, (5.17)
curvature drift : v
c
=
p

qB
2
B(B grad)B
B
2
, (5.18)
polarization drift : v
P
=
m
q

B
2
. (5.19)
A physical explanation of the electric drift (5.14) is given above. The gravitational drift (5.15)
may be derived from the electric drift by replacing qE in (5.1) by mg, and thence in (5.14) to
obtain (5.15). Note that the drift is perpendicular to both the gravitational eld and to the
magnetic eld. Also, it is in opposite senses for charges of opposite signs. Charges of opposite
signs owing in opposite directions imply an electric current. In a uid description, the current
density J implies a force per unit volume J B that opposes the gravitational force density g
on the uid of mass density . The fact that the force due to the current opposes the initial force
that drives the current is an example of Lenz law. Physically, the gravitational drift may be
understood as illustrated in Figure 5.1. The gravitational force accelerates particles downward, so
that they have higher perpendicular momenta near the bottom of their orbits, and so, according
to (5.4), have larger gyroradii there than near the top of their orbits.
5.3 Drift motions 43
Figure 5.1: The gravitational drift of a positively-charged particle is illustrated for the case where
the magnetic eld is into the page and the gravitational force is directed downward.
The gradient drift (5.16) is associated with a change in the strength of the magnetic eld. The
drift is perpendicular to both the direction of the magnetic eld and to the direction of grad B,
and is opposite for oppositely charged particles. In this case the implied current generates a
magnetic eld that is such as to oppose the gradient in B. The eld generated by the induced
current satises curl B =
0
J which is Amp`eres Law in the case of constant or zero E. Physically,
the gradient drift may be understood in terms of an argument similar to that used to explain the
gravitational drift. As illustrated in Figure 5.2, if one takes an idealized case in which the magnetic
eld changes abruptly at a surface that passes through the center of gyration of the particle, then
the gyroradius is dierent in two halves of the orbit. On joining a sequence of semicircles with
radii that alternate between two values, one obtains the orbit illustrated in Figure 5.2, which
shows that a drift motion results.
The inertial drift (5.17) is attributed to the coordinate frame in which the spiraling motion is
described not being an inertial frame. For example, in a rotating plasma, the frame in which the
plasma is momentarily at rest is not an inertial frame. The quantity (dp/dt)
0
, which is the time
derivative of the momentum relative to an inertial frame (e.g., the instantaneous rest frame), is
the inertial force.
The curvature drift (5.18) and the polarization drift (5.19) are both specic examples of in-
ertial drift. Curvature (or centrifugal) drift is associated with curvature of the magnetic eld.
Introducing the unit vector b = B/B, one has
B(B grad)B
B
3
= b (b gradb), b gradb =
n
R
c
, (5.20)
where R
c
is the radius of curvature of the eld lines, and where n is a unit vector along the direction
44 5. Orbit theory
(a)
(b)
Figure 5.2: (a) An idealized case which demonstrates how the gradient drift occurs: the magnetic
eld is into the page and its strength increases abruptly at a surface (dashed line) that passes
through the gyrocenter of the negatively-charged particle so that the gyroradii are dierent on the
two sides. The curve is drawn by joining semicircles with radii r
1
and r
2
.
(b) A curved magnetic eld line may be approximated by the arc of a circle: the radius of curvature
R
c
is the radius of this circle.
toward the center of gyration, as illustrated in Figure 5.2b. Polarization drift is associated with a
time-varying electric eld, in which case the inertial drift is (dp/dt)
0
= mdv
E
/dt.
An alternative way of writing the drift motions is in terms of the average (over the spiraling
motion) position R = X, with X given by (5.6) in the case of a uniform eld. Retaining only
the electric, gradient and curvature drift one has

R = v

b +
E b
B
+
v

2qB
_
b gradB
B
_
+
v

2qB
b (b grad )b. (5.21)
It is not at all obvious but it can be shown that the equation of motion for the gyrocenter has the
following components
p

= qb E +
1
2
v

div b, p

=
1
2
v

div b. (5.22)
Magnetic elds do no work (since the force is always perpendicular to the motion) so although
the magnetic eld can induce drifts we expect any change in the particle energy to be determined
solely by the electric eld. Moreover, the perpendicular component of the electric eld can be
removed by a Lorentz transformation, and so in the simplest approximation it too does no work.
5.4 Adiabatic invariant: magnetic trapping 45
If one writes = (p
2

+p
2

)/(2m) then = v

+v

and the substitution of (5.22) leads to the


expected result = qE

.
5.4 Adiabatic invariant: magnetic trapping
Consider a charge spiraling in a magnetic eld, B, that is changing slowly (compared to a gyro-
radius) in space. The motion is periodic is the angle known as the gyrophase, which is equal
to
0
+ t in the case (5.6) of a uniform eld. The conserved quantity is the angular momentum
associated with this periodic motion, that is the -component of p times the radius, R, of gyra-
tion. The -component of p is equal to p

and this gives an adiabatic invariant 2p

R, where
the 2 arises in a more formal denition, given by the integral in (5.27). Since R p

/B, this
contribution is proportional to p
2

/B. Hence, one nds that


p
2

B
= constant (5.23)
is an adiabatic invariant, sometimes called the rst adiabatic invariant, and sometimes referred to
an the magnetic moment of the particle. (A charge moving in a circle corresponds to a current
loop and the magnetic moment is that associated with this current loop.)
The adiabatic invariant (5.25) may be shown to be an invariant using the result (5.22) from
the theory of drift motions. First note the following result:
d
dt
1
B
= v

b grad
1
B
=
v

div b
B
, (5.24)
where the rst identity follows for B/t = 0, and where b = B/B and div B = 0 are used in the
second identity. Then using (5.22) and (5.24), one nds
d
dt
_
p
2

B
_
=
2p

B
+p
2

d
dt
1
B
= 0. (5.25)
One implication of the conservation of the rst adiabatic invariant is the reection of a particle
from a magnetic compression. In the absence of any eld other than an inhomogeneous magnetic
eld, one has p = constant, and hence (5.25) implies sin
2
/B = constant, where is the pitch
angle of the particle, cf. (5.3). Hence, as a particle propagates in a direction of increasing bgradB,
sin
2
increases B. As sin
2
increases, [ cos [ decreases and so [p

[ decreases. If sin
2
reaches
unity then the particle motion is strictly circular, with p

= 0, and the particle reects at that


46 5. Orbit theory
Figure 5.3: The motion of a trapped particle inside a magnetic bottle is illustrated schematically.
point and moves back in the direction of decreasing b gradB. This leads to the concept of a
magnetic bottle, which is a region of weak B between regions of stronger B such that particle
can reect at either end, as illustrated in Figure 5.3. Note that particles with suciently small
in the center of the bottle are not reected, and escape from the ends of the bottle. The range
<
0
for which particles are not trapped is called the loss cone.
5.5 Appendix: formal theory of adiabatic invariants
Any mechanical system that has one or more nearly periodic motions has a nearly conserved
quantity corresponding to each such motion; these conserved quantities are called adiabatic in-
variants. Formally, this may be seen simply in terms of Lagrangian or Hamiltonian dynamics. In
terms of Lagrangian dynamics, let us choose one of the generalized coordinates to be the angle,
, corresponding to the quasiperiodic motion, so that the Lagrangian for the system is L(,

),
where the dependence on other variables is of no interest. The Lagrangian equation of motion is
d
dt
_
L

= 0. (5.26)
Suppose one integrates (5.26) over one period of the motion, say over 0 < < 2. The congu-
ration of the system is the same at = 2 as at = 0 so that the nal term in (5.26) integrates
to zero. Thus (5.26) implies that the time derivative of a quantity is zero and hence that the
quantity is conserved. Thus one nds
_
d
L

= constant,
_
dQP = constant, (5.27)
which is the desired adiabatic invariant. The second form in (5.27) is the corresponding form in
Hamiltonian dynamics, when the periodic motion is in an arbitrary generalized coordinate Q with
5.5 Appendix: formal theory of adiabatic invariants 47
conjugate momentum P.
A subtle point that is that the conjugate 3-momentum (conjugate to x) is p + qA, where A
is the vector potential of the magnetic eld. This is not important in the derivation of the rst
adiabatic invariant (5.23), but it is important when considering the full set of adiabatic invariants
for a particle in a dipolar-like eld. There are three adiabatic invariants, often denoted M, J, .
M is the invariant (5.23).
A particle trapped in a magnetic bottle has a quasiperiodic motion corresponding to its bounce
motion between the reection points. There is an adiabatic invariant corresponding to this motion,
sometimes called the second adiabatic invariant. Let the distance, s, along the eld lines be a
generalized coordinate, whose conjugate momentum is p

. Then (5.27) implies


J =
_
ds p

= constant, (5.28)
where the integral is along the orbit of the gyrocenter between the reection points.
There is a third adiabatic invariant for particles trapped in a magnetic eld which is roughly
dipolar, as is the case for the Earths magnetic eld within several Earth radii, R
E
. The curvature
drift causes particles to drift (in magnetic longitude) around the Earth. As this drift is quasiperi-
odic there is an adiabatic invariant associated with it. This invariant is given by the integral of
the component of qA in the direction of the drift around the closed orbit, which integral involves
the radial distance r = LR
E
, implying that the orbit (rather the center of the bounce motion) of
the particle is conned to a given r or, as is standard jargon in this context, to a given L shell.
48 5. Orbit theory
5.6 Exercise Set 5
5.1 Calculate the gyroradius of a particle under the following conditions.
(a) A 1 eV electron with pitch angle = /2 in a laboratory devise where the only magnetic eld
is that of the Earth, assuming B = 0.3 G.
(b) A 2 keV electron with pitch angle = /2 at the Earths magnetic equator at L = 3. Assume
that the Earths magnetic eld is dipolar with the magnetic eld at the pole B = 0.3 G. Express
your answer in centimeters.
(c) A 10
18
eV ion with pitch angle = /2 in the interstellar magnetic eld with B = 3 G.
Express your answer in parsecs (1 pc = 3 10
16
m).
5.2 An MHD generator is a dynamo that converts mechanical energy into electrical energy. The
mechanical energy is in the form of a partially ionized gas forced (blown by a fan for example)
across a magnetic eld. The ow across the magnetic eld creates an electric eld which is such
that the electric drift is equal to the ow velocity. This electric eld is due to a (forced) charge
separation in the plasma. If one puts conducting plates on either side of the ow, surface charges
of opposite sign collect on the two plates. The dynamo operates when one connects the two plates
by a wire (outside the plasma). The voltage associated with the dynamo is found by integrating
the electric eld along a line between the two plates.
Let the ow velocity, u, be along the x axis, and the magnetic eld, B, be along the z axis.
The electric eld is along the y axis.
(a) Derive a formula for the electric eld. (b) Derive a formula for the voltage assuming the plates
are a distance L apart. (c) Estimate the voltage for a ow u = 1 ms
1
between plates L = 1 m
apart in a magnetic eld B = 1 T.
5.3 The magnetic eld lines associated with a line current, I, are circles around the axis dened
by the current line. The azimuthal component of the magnetic eld is B

=
0
I/4r, where r is
the radial distance from the axis. Consider a particle with gyroradius much smaller than r moving
around the circular eld line at r, such that its center of gyration has a velocity v.
Find expressions for
(a) the gradient drift, and
(b) the curvature drift.
5.4 Assume the Earths magnetic eld is a dipole, implying that B in the equatorial plane decreases
5.6 Exercise Set 5 49
1/r
3
, with the radial distance r = LR
E
, where R
E
= 6.4 10
6
m is the radius of the Earth.
Particles trapped in the eld drift in azimuthal angle around the Earth due to the grad-B drift.
Electrons and ions drift in opposite sense, and their relative drift implies a current. The drift of
particles trapped in the Earth so-called radiation or van Allen belts produce a ring current.
Estimate the time it takes for a particle to drift around the Earth at L = 4 for
(a) a 1 keV electron, and
(b) a 1 MeV ion.
5.5 A plasma consists of electrons and protons with equal number densities, n, with a uniform
magnetic eld along the x-axis with z the vertical direction.
(a) Calculate the current density, J, due to the gravitational drift of electrons and protons, where
the gravitational acceleration, g, is along the negative z-axis.
(b) Calculate the force per unit volume, J B, due to this current density.
(c) Give a physical interpretation of your answer.
50 5. Orbit theory
Chapter 6
Cold magnetized plasma
In the absence of a magnetic eld, the response of a cold plasma can be described by a dielectric
constant K() = 1
2
p
/
2
. The dispersion equation implies longitudinal waves at K() = 0,
implying oscillations at =
p
, and transverse waves at n
2
= K(). When a magnetic eld is
included, the response of a cold plasma is anisotropic, and needs to be described by a tensor.
In this lecture derive this tensor for a cold plasma consisting of electrons and various species of
positive ions. The dispersion equation becomes a quadratic equation for n
2
implying that there
are two dierent natural wave modes of a cold plasma. At high frequencies these become the
magnetoionic waves, and at low frequencies they are eectively the MHD modes for zero sound
speed. These limiting cases are discussed in later lectures. In this lecture we are concerned
primarily with describing the response of a cold plasma, and the procedure for calculating the
properties of the natural modes of the anisotropic medium.
6.1 Response of a cold plasma
The response for a cold magnetized plasma may be found by solving the equation of motion for
particles of species , with mass m

and charge q

:
m

dv
dt
= q

(E +v B). (6.1)
52 6. Cold magnetized plasma
Expanding in plane waves and assuming that B is along the z axis, (6.1) may be written in matrix
form:
i
_
_
_
_
_
_
_
v
x
v
y
v
z
_
_
_
_
_
_
_
=
q

_
_
_
_
_
_
_
E
x
E
y
E
z
_
_
_
_
_
_
_
+
q

B
m

_
_
_
_
_
_
_
v
y
v
x
0
_
_
_
_
_
_
_
. (6.2)
A rearrangement gives
i
_
_
_
_
_
_
_
i

0
i

0
0 0
_
_
_
_
_
_
_
_
_
_
_
_
_
_
v
x
v
y
v
z
_
_
_
_
_
_
_
=
q

_
_
_
_
_
_
_
E
x
E
y
E
z
_
_
_
_
_
_
_
, (6.3)
with

= q

/[q

[,

= [q

[B/m

. Solving the matrix equation gives


_
_
_
_
_
_
_
v
x
v
y
v
z
_
_
_
_
_
_
_
=
i

_
_
_
_
_
_
_

2
i

0
i


2
0
0 0
2

_
_
_
_
_
_
_
_
_
_
_
_
_
_
E
x
E
y
E
z
_
_
_
_
_
_
_
. (6.4)
The current density for species is J

= q

v. After summing over the contributions of all


species (electrons and ions), the current may be used to identify the dielectric tensor.
Cold plasma dielectric tensor
The relation between the induced current and the electric eld denes the conductivity tensor,
() say. The contribution of species to () follows by multiplying (6.4) by q

. After
summing over species this gives
() =

q
2

_
_
_
_
_
_
_

2
i

0
i


2
0
0 0
2

_
_
_
_
_
_
_
. (6.5)
The current may be written in the form J = P/t, and the relation between P and E denes
the susceptibility tensor, () = i()
0
. The dielectric tensor is identied as the unit tensor
plus ().
6.2 Dispersion equation for a cold plasma 53
(b)
z
o
x
(a)
n
2
2 1.5 0.5
1
-2
o
x
p
/
Figure 6.1: Refractive index curves for the magnetoionic waves for
e
/
p
= 0.5. (a) For = 0
there are two curves plus a vertical line (not shown) at =
p
. (b) Circled portion of (a)
magnied; the dashed lines is for ,= 0. For = 0 the o mode and the z mode join at =
p
A standard form for the resulting expression for the dielectric tensor for a cold plasma is
1
K() =
_
_
_
_
_
_
_
S() iD() 0
iD() S() 0
0 0 P()
_
_
_
_
_
_
_
, (6.6)
S() =
1
2
[R
+
() + R

()], D() =
1
2
[R
+
() R

()],
R

() = 1

2
p

, P() = 1

2
p

2
, (6.7)
where the sum is over species, with the th species having mass m

, charge q

[q

[, number
density n

, plasma frequency
p
= (q
2

/
0
m

)
1/2
.
6.2 Dispersion equation for a cold plasma
The wave equation can be written in the matrix form (4.6), that is, as E = 0, with =
n
2
[ 1] +K. The matrix form for , with the coordinate axes chosen such that B is along the
z axis and is in the x-z plane at an angle to B, is
=
_
_
_
_
_
_
_
S n
2
cos
2
iD n
2
sin cos
iD S n
2
0
n
2
sin cos 0 P n
2
sin
2

_
_
_
_
_
_
_
, (6.8)
1
T.H. Stix Waves in Plasmas, McGraw-Hill (1962)
54 6. Cold magnetized plasma
The dispersion equation is found by setting the determinant of this matrix to zero. This gives a
quadratic equation for n
2
:
[[ = An
4
Bn
2
+C = 0, (6.9)
with
A = S sin
2
+P cos
2
, B = (S
2
D
2
) sin
2
+PS(1 + cos
2
), C = P(S
2
D
2
). (6.10)
The solutions may be written in the form
n
2
= n
2

=
B F
2A
, F = (B
2
4AC)
1/2
. (6.11)
The two solutions dene two modes. However, these correspond to propagating waves only for
n
2
> 0. For n
2
< 0 the solutions are said to describe evanescent waves: solutions that oscillate in
time by decay exponentially in space.
6.3 Polarization vectors
The polarization vector e
M
(k) for any wave mode M in a magnetized plasma may be expressed
in terms of the set of basis vectors
= (sin , 0, cos ), t = (cos , 0, sin ), a = (0, 1, 0). (6.12)
These are unit vectors along the wave vector k, along the direction perpendicular to k in the Bk
plane, and along the direction orthogonal to both B and k, respectively. The component of the
electric vector along a is out of phase with the components in the Bk plane. It is convenient to
write
e
M
=
L
M
+T
M
t +ia
(L
2
M
+T
2
M
+ 1)
1/2
, (6.13)
with M = for the cold plasma modes, and M = o, x for the magnetoionic waves. The longitu-
dinal part of the polarization vector is described by L
M
and the transverse part is described by
T
M
.
The transverse part corresponds to an elliptical polarization, with [T
M
[ the axial ratio of the
polarization ellipse, as illustrated in Figure 6.2. By denition, [T
M
[ is the ratio of the moduli
of the component along t to the component along a. The triad of unit vectors (6.12) forms a
right hand set, and hence the sign of T
M
determines the handedness of the ellipse, with T
M
> 0
corresponding to right hand polarization and T
M
< 0 corresponding to left hand polarization.
6.4 Polarization of cold plasma waves 55

D
O
B
A
C
Figure 6.2: The axial ratio [T[ is equal to [AC[/[BD[. The solid ellipse corresponds to T > 0.
The wave is propagating into the page. The two ellipses correspond to orthogonal polarizations.
6.4 Polarization of cold plasma waves
So far we have only considered the condition for a solution of the wave equation to exist. When
this condition is satised, a solution of the matrix equation for E exists. The amplitude and phase
of the solution are arbitrary. It is convenient to choose them such that the solution corresponds
to a polarization vector of the form (6.13). This involves solving for the parameters T
M
, L
M
, with
M = here.
The polarization vectors are constructed from any column of the matrix of cofactors of . One
choice gives
T
M
=
DP cos
An
2
M
PS
, L
M
=
(P n
2
M
)Dsin
An
2
M
PS
. (6.14)
On inserting explicit expressions for the refractive indices into (6.14) one nds explicit expressions
for the polarization vectors. However, for computational and other purposes it is more convenient
to note that there is a linear relation between n
2
and 1/T, and that because n
2
satises a quadratic
56 6. Cold magnetized plasma
equation, 1/T and T must also satisfy quadratic equations. For T this equation is
T
2

(PS S
2
+D
2
) sin
2

PDcos
T 1 = 0. (6.15)
It is straightforward to solve the quadratic equation (6.15) for T = T

and calculate n
2

and L

in terms of T

by inverting (6.14). This allows one to make approximations systematically: one


approximates T

, and evaluates the corresponding approximations to n


2

and L

using (6.14).
6.5 Exercise Set 6 57
6.5 Exercise Set 6
6.1 At very low frequencies, the two cold modes reduce to the Alfven and magnetoacoustic waves.
(a) Show that in the limit 0 (6.6) with (6.7) implies S(0) = 1 +

2
p
/
2

, D(0) = 0,
P(0) = .
(b) Show that in this limit the dispersion relations for the two modes reduce to n
2
= S(0)/ cos
2
,
n
2
= S(0).
(c) Show that one has S(0) = 1 + c
2
/v
2
A
.
(d) Hence show that, for v
2
A
c
2
, the two dispersion relations become
2
= k
2
v
2
A
cos
2
,
2
=
k
2
v
2
A
, respectively.
6.2 Consider the response of a charge-neutral cold plasma at low frequencies. The plasma is
assumed to be composed of electrons and various species of positive ions with charge q
i
= Z
i
e,
mass m
i
= A
i
m
proton
and number density n
i
.
(a) Show that the charge neutrality condition n
e
=

i
Z
i
n
i
implies

2
p

e
=

2
pi

i
. (6.16)
(b) With the Alfven speed dened by v
A
= B/(
M
)
1/2
, where
M
is the mass density, show that
if the mass of an electron is neglected compared to that of an ion, then one has

2
pi

2
i
=
c
2
v
2
A
. (6.17)
58 6. Cold magnetized plasma
Chapter 7
Magnetoionic theory
Historically, the magnetoionic theory
1
was the second important contribution (after Langmuirs
work) to what developed into modern plasma physics. The motivation was to understand radio
wave propagation in the ionosphere. In this application, the Earths magnetic eld plays an
important role, and the thermal motions of the electrons are unimportant. This corresponds
to a cold magnetized electron gas. The ions play no role, and the name magnetoionic is an
anachronism. The magnetoionic waves are important in understanding wave propagation in radio
astronomy.
7.1 Magnetoionic parameters
In the magnetoionic theory only the contribution of the electrons is retained. The plasma fre-
quency,
p
= (e
2
n
e
/
0
m
e
)
1/2
, the electron cyclotron frequency,
e
= eB/m
e
, and the wave fre-
quency, , are combined into two magnetoionic parameters:
X =

2
p

2
, Y =

e

. (7.1)
The dielectric tensor (6.6) has components
S =
1 X Y
2
1 Y
2
, D =
XY
1 Y
2
, P = 1 X. (7.2)
The coecient (6.10) become
A = [1 X Y
2
+XY
2
cos
2
]/(1 Y
2
),
1
developed by Appleton and Hartree in the early 1930s
60 7. Magnetoionic theory
(b)
z
o
x
(a)
n
2
2 1.5 0.5
1
-2
o
x
p
/
Figure 7.1: Refractive index curves for the magnetoionic waves for
e
/
p
= 0.5. (a) For = 0
there are two curves plus a vertical line (not shown) at =
p
. (b) Circled portion of (a)
magnied; the dashed lines is for ,= 0. For = 0 the o mode and the z mode join at =
p
B = [2(1 X)
2
2Y
2
+XY
2
(1 + cos
2
)]/(1 Y
2
),
C = (1 X)[(1 X)
2
Y
2
]/(1 Y
2
). (7.3)
For the magnetoionic modes, the two solutions (6.11) can be rewritten as
n
2

= 1
X(1 X)
1 X
1
2
Y
2
sin
2
+
, (7.4)
with = 1 and with

2
=
1
4
Y
4
sin
4

2
+ (1 X)
2
Y
2
cos
2
. (7.5)
The two solutions are called the ordinary (o) and extraordinary (x) modes. The technical denition
of the ordinary mode is that it is the mode that has n
2
1X for /2; for >
p
(X < 1),
this denition corresponds to n
2
o
= n
2
+
and n
2
x
= n
2

.
7.2 Cuto frequencies
Transverse waves in an isotropic plasma have n
2
= 1
2
p
/
2
, and so they exist as propagating
waves only for >
p
, where n
2
is positive. The frequency where n
2
becomes zero is referred to
as the cuto frequency.
The cuto frequencies for the o- and x-modes are at n
2
+
= 0 and n
2

= 0, respectively. More
generally, cutos occur at n
2
= 0, where (6.9) implies C = 0, which gives
P(S
2
D
2
) = 0. (7.6)
7.3 High-frequency limit 61
n
2
o
z
x
2 1
-2
0
2
p
/
Figure 7.2: As in Figure 7.1 but for = 30

.
There are three solutions of (7.6) for the magnetoionic waves. One corresponds to P = 0, which
gives =
p
. This cuto is in the o mode. There are two positive frequency solutions of
S
2
D
2
= 0. These are at =
x
, and =
z
, with

x
=
1
2

e
+
1
2
(4
2
p
+
2
e
)
1/2
,
z
=
1
2

e
+
1
2
(4
2
p
+
2
e
)
1/2
. (7.7)
The cuto
x
applies to the x mode and that at
z
applies to the z mode, which is not discussed
further here.
7.3 High-frequency limit
In the high-frequency limit,
p
,
e
, the properties of the magnetoionic modes simplify.
The simplest way to treat this limit is to make the approximation that the modes are circularly
polarized, which is valid except for a small range of angles about propagation perpendicular to
the magnetic eld lines. In this approximation, the refractive indices are given by
n
2

= 1
X
1 +Y [ cos [
, (7.8)
with = 1 for the o mode and = 1 for the x mode. On expanding in X 1 and Y 1,
(7.8) gives
n

1
1
2
X +
1
2
XY [ cos [. (7.9)
The handedness of the polarization is dened in terms of a screw sense relative to the direction
of the magnetic eld lines.
62 7. Magnetoionic theory
1
2
-10
10
20
0
w
z
x
o
n
2
p
/
Figure 7.3: As for Figure 7.1 but plotted on a dierent scale. The whistler (w) mode branch is in
the upper left hand corner.
The handedness of observed radio emission is dened as a screw sense relative to the direction,
, of wave propagation. This is the same as the screw sense relative to b for < /2 and the
opposite sense for > /2. For some purposes, it is convenient to label the refractive index in
terms of right (r) and left (l) hand polarizations:
n
r,l
1
1
2
X
1
2
XY cos . (7.10)
The dierence between the refractive indices causes Faraday rotation, which is important in
radio astronomy. Faraday rotation is the rotation of the plane of linear polarization as radiation
propagates through a magnetized medium whose waves modes are circularly polarized. One can
understand Faraday rotation qualitatively from the following idealized example. Suppose radiation
at its source, at s = 0, is linearly polarized along the 1-axis, and that it propagates along the 3-axis.
One can separate the initial linear component into right and left hand circularly components of
equal amplitude. After the radiation has propagated a distance s, the refractive index dierence,
n say, implies that the two modes are out of phase with each other by k s, with k = n/c
the dierence in wavenumber. On recombining the two circularly polarized components, they give
a linearly polarized component whose plane of polarization is rotated from its original direction.
The plane of linear polarization rotates at a rate k/2 per unit length.
7.4 Polarization of magnetoionic waves 63
7.4 Polarization of magnetoionic waves
The polarization vectors for the magnetoionic waves are constructed from any column of the
matrix of cofactors of . One choice gives (6.14). An alternative procedure is to solve (6.15) for
T = T

and use (6.14) to nd n


2
and L in terms of T. For the magnetoionic modes (6.15) becomes
T
2
+
Y sin
2

(1 X) cos
T 1 = 0. (7.11)
The solutions of (7.11) are
T = T

=
Y (1 X) cos
1
2
Y
2
sin
2

=

1
2
Y
2
sin
2

Y (1 X) cos
, (7.12)
with
2
given by (7.4), and where = 1 corresponds to the o mode and = 1 to the x mode.
The two polarization ellipses are orthogonal in the sense
T
+
T

= 1. (7.13)
Approximations to the axial ratio follow by considering the ratio of the two terms in the square
root for F, cf. (6.11) or , cf. (7.4). For [(1 X) cos [
1
2
Y sin
2
one nds
T


cos
[ cos [
1 X
[1 X[
_
1 +
Y sin
2

2[(1 X) cos [
+
_
. (7.14)
The leading terms in (7.14) correspond to circular polarization, with the handedness such that the
electric vector in the x mode and in the whistler mode rotate in the same sense as that in which
electrons gyrate (right hand screw sense relative to B) and the electric vector in the o mode and
the z mode rotate in the opposite sense. This is called the quasi-circular limit. The corresponding
approximation to the dispersion relations is, for Y [ cos [ 1, and X < 1,
n
2

= 1 X(1 Y [ cos [ + ) = 1

2
p

2
(1

e
[ cos [

+ ). (7.15)
In the opposite limit [(1X) cos [
1
2
Y sin
2
, called the quasi-planar limit or quasi-linear limit,
one has
T
o
, n
2
o
1 X, L
o

XY sin
1 X
; (7.16)
T
x
0, n
2
x
1
X(1 X)
1 X Y
2
+XY
2
cos
2

,
L
x

XY sin
1 X Y
2
+XY
2
cos
2

. (7.17)
The transverse parts of the polarization correspond to linear polarizations along t and a for the
ordinary and extraordinary modes respectively.
64 7. Magnetoionic theory
7.5 Exercise Set 7
7.1 The low-frequency branch of the magnetoionic o-mode is called the whistler mode, where the
name originates from whistling atmospherics heard in early radio receivers. (These waves are
often called helicon waves in laboratory applications.) This exercise is to derive the properties of
the whistler mode.
(a) Show that in a plasma with
2
e

2
p
, (7.5) implies = [(1 X)Y cos [ except for a small
range of angles about = /2.
(b) Show that for
2

2
e

2
p
the solution = +1 in (7.4) implies
n
2
o

2
p

e
[ cos [
. (7.18)
(c) Show that the polarization vector for the whistler modes in this approximation is
e
o
=
(1, i[ cos [, 0)
(1 + cos
2
)
1/2
. (7.19)
(d) Evaluate the partial derivatives in the expression
v
go
=
c
(n
o
)/
_

1
n
o
n
o

t
_
(7.20)
for the group velocity.
(e) Hence show that the group velocity is
v
go
=
c
n
o
(sin , 0, cos + sec ). (7.21)
(e) Dene the ray angle by writing v
go
= [v
go
[(sin
r
, 0, cos
r
), show that one has
cos
2

r
=
(1 + cos
2
)
2
1 + 3 cos
2

. (7.22)
(f) Show that there is maximum ray angle sin
r
= 1/3 corresponding to sin
2
= 2/3.
7.2 A formal treatment of Faraday rotation involves the Stokes parameters, I, Q, U, V , which
involve the outer produce of the wave amplitude and its complex conjugate. The degrees of
polarization are
p =
(Q
2
+U
2
+V
2
)
1/2
I
, p
l
=
(Q
2
+U
2
)
1/2
I
, p
c
=
V
I
. (7.23)
7.5 Exercise Set 7 65
The axial ratio, T, and the angle that denes the plane of linear polarization are determined by
Q
pI
=
_
T
2
1
T
2
+ 1
_
cos 2,
U
pI
=
_
T
2
1
T
2
+ 1
_
sin 2,
V
pI
=
2T
T
2
+ 1
. (7.24)
Show that the rate of Faraday rotation is given by
d
ds
=

2
p

e
cos
2c
2
(7.25)
where (7.10) is used.
7.3 The plane of polarization is an observable quantity, and measurement of it at several frequen-
cies provides information of the properties of the medium along the ray path between the source
and the telescope. Assuming a homogeneous medium along the ray path, the angle through which
the plane of polarization is rotated is
=

2
p

e
2
2
c
Dcos , (7.26)
where D is the distance to the source. Taking variations in the properties of the medium along
the ray path into account, (7.26) is replaced by an integral along the ray path:
=
e
3
2
0
m
2

2
c
_
D
0
ds Bn
e
. (7.27)
The dependence on frequency implies a dependence on the square of the wavelength, = 2c/.
If one measures at dierent wavelengths one can determine the constant of proportionality,
which is called the rotation measure (RM).
Show that the rotation measure, dened by writing = RM
2
, is given by
RM =
e
3
2(2)
2

0
m
2
c
2
_
D
0
ds Bn
e
. (7.28)
The units of RM are inverse length square, usually m
2
.
7.4 The polarization of transverse waves is described by the matrix
p =
1
2
_
_
_
1 +p
Q
p
U
ip
V
p
U
+ip
V
1 p
Q
_
_
_. (7.29)
The degree of polarization, p, may be identied by writing p =
1
2
(1 p)1 +p ee

with
p = (p
2
Q
+p
2
U
+p
2
V
)
1/2
. (7.30)
66 7. Magnetoionic theory
If two signals with intensities I
1
and I
2
and polarization matrices p
1
and p
2
are added incoherently,
the polarization matrix of the resulting radiation is
p =
I
1
p
1
+I
2
p
2
I
1
+I
2
. (7.31)
Consider two signals of equal intensity being added, one that is completely linearly polarized
(p
Q
= 1) and the other that is completely circularly polarized (p
V
= 1).
(a) What is the degree of polarization of the combined radiation?
(b) What is the axial ratio of the polarization ellipse?
(c) How does the result change if you assume p
U
= 1, rather than p
Q
= 1, for the linearly
polarized component?
7.5 The magnetoionic waves are nearly circularly polarized and nearly linearly polarized in two
opposite limits depending on which term dominates in the expression (7.5) for
2
.
(a) Show that the angle at which the two contributions are equal is

0
= arcsin
_
2
1/2
Y
[(1 X)
4
+Y
2
(1 X)
2
]
1/2
(1 X)
2

1/2
_
. (7.32)
(b) Show that (C.2) reduces to
0
arccos
1
2
Y for X 1, Y 1.
(c) Estimate the range of angles for which the approximation of circular polarization breaks down
for 1 GHz radiation in the fully ionized region of the ISM.
7.6 Consider the response of a charge-neutral cold plasma at low frequencies. The plasma is
assumed to be composed of electrons and various species of positive ions with charge q
i
= Z
i
e,
mass m
i
= A
i
m
proton
and number density n
i
.
(a) Show that the charge neutrality condition n
e
=

i
Z
i
n
i
implies

2
p

e
=

2
pi

i
. (7.33)
7.5 Exercise Set 7 67
(b) With the Alfven speed dened by v
A
= B/(
0
)
1/2
, where is the mass density, show that if
the mass of an electron is neglected compared to that of an ion, then one has

2
pi

2
i
=
c
2
v
2
A
. (7.34)
7.5 Double solutions of the dispersion equation occur when both and its derivative, for example,
with respect to n
2
, vanish simultaneously.
(a) Show that a double solution of the dispersion equation for the magnetoionic waves occurs at
X = 1, sin = 0.
(b) Plot n
2
as a function of X 1 for small sin .
7.8 Show that the refractive index curves for the magnetoionic modes pass through the points
n
2
= 1 and n
2
= 0 at =
p
, and determine the polarization vectors at these two points.
68 7. Magnetoionic theory
Chapter 8
Magnetohydrodynamics
The most general theoretical description of a plasma is in terms of kinetic theory. However, for
many purposes a simpler description in terms of a uid theory suces. The uid theory for
a magnetized plasma is magnetohydrodynamics, usually abbreviated to MHD. Here the MHD
equations are written down, and some simple applications are discussed briey.
8.1 MHD equations
A ow of a uid is described by uid velocity, denote u here. It is important not to confuse the
uid velocity with particle velocities. The uid velocity of the air in this room involves relatively
slow motions, very much slower that the speed of individual molecules. An element of the uid
consists of an average number of molecules, but individual molecules are entering and leaving the
uid element very rapidly. In a uid description, the uid velocity is a function of position and
time, u(t, x). The uid velocity is the time derivative of the uid displacement, denoted (t, x).
The uid equations can be derived from kinetic theory, by taking moments of the distribution
function and of the Boltzmann equation. However, it is usual to simply write them down using
arguments based on uid mechanics.
Let (t, x) be the mass density of the uid. The mass of the uid is a conserved quantity, and
this implies that an equation of mass continuity is satised. This equation is

t
+ div (u) = 0. (8.1)
The dependence of and u on t, x is implicit in (8.1). One can see that (8.1) described conservation
70 8. Magnetohydrodynamics
of mass by integrating over a volume, V . The integral of over the volume gives the mass of
the uid inside the volume. The integral of the rst term in (8.1) gives the rate of change of
the mass in the volume with time. The volume-integral of the divergence of a vector eld may
be written as a surface integral over the normal of the vector eld, using Stokes theorem. The
integral of the second term in (8.1) gives the rate of ow of mass across the surface of the volume
V . Conservation of mass requires that the sum of these two terms be zero for any volume V , and
(8.1) ensures that this condition is satised.
The equation of motion for the uid is

_
u
t
+ (u grad )u
_
= grad P +E +J B+f. (8.2)
The left hand side is the mass density time the uid acceleration, du/dt. The uid velocity, u(t, x),
is an Eulerian velocity, and its total time-derivative consists of two parts. The term u/t is due
to the explicit dependence of u(t, x) on t. The second term on the left hand side is called the
convective derivative. It takes into account the change in the vector eld u(t, x) with position
along the ow lines, The right hand side of (8.2) is the force per unit volume. The pressure
gradient, gradP, is a uid force that has no obvious counterpart in conventional mechanics. The
next two terms on the right hand side arise from the force q(E+vB) on each charge in the uid.
On summing over all the charges, the charge density is =

, and the current density is


J =

, where u

is the uid velocity of species . The nal term in (8.2) allows for a
mechanical force, f, per unit mass; one has f g when the force is gravity, with g the acceleration
due to gravity. In most applications of MHD the charge density is assumed to be zero, and the
term E is omitted.
The electrical conductivity, of the uid leads to a relation between the current density and
the eective electric eld. This is
J = (E +u B), (8.3)
where E + u B is equal to the electric eld in a frame in which the uid is at rest. For many
purposes, (8.3) can be simplied to E+uB = 0 by assuming = . The perfectly-conducting
limit corresponds to the absence of resistive eects, and corresponds to ideal MHD. When a nite
value of is included one refers to resistive MHD.
In addition to these equations, the MHD equations include two further sets of equations. One
8.2 Ideal MHD 71
is Maxwells equations
curl E = B/t, curl B =
0
J + (1/c
2
) E/t,
div E = /
0
, div B = 0, (8.4)
with the displacement current usually omitted from (8.4).
An additional equation is required, and this is an equation of state for the uid. Three
alternative equations of state are used in dierent contexts: the incompressible, isothermal and
adiabatic equations of state. The incompressible conditions corresponds to = constant, and then
(8.1) implies div u = 0. The isothermal equation of state is T = constant. The adiabatic equation
is P

, with the adiabatic index, which has the specic value = 5/3 for a monatomic gas.
Thus one has
div u = 0, T = constant, P

, (8.5)
for the incompressible, isothermal and adiabatic cases, respectively. The incompressible equation
of state (with grad P = 0) is appropriate when MHD theory is applied to a conducting liquid,
such as a column of mercury. It is sometimes useful for gaseous plasmas, specically when the
eects of interest are known to involve no compressions, as in an Alfven wave. The isothermal
equation of state is appropriate for processes that occur on such a long time scale that there is
time for thermal conduction to maintain a constant temperature. This condition is not satised
in most applications of interest. The adiabatic equation (8.5) applies to compressible MHD.
8.2 Ideal MHD
The condition for ideal MHD is E+ u B = 0, which follows from (8.3) for . Taking the
components parallel and perpendicular to the magnetic eld implies
E

= 0, E

= u

B. (8.6)
An interpretation of (8.6) follows from the electric eld drift (5.14) due to E

, which is v
E
= u

.
The electric eld (8.6) is that required to cause all particles in the plasma to drift across the
magnetic eld lines at u

.
An important feature of ideal MHD is that the magnetic eld and the plasma move together.
One says that the magnetic eld is frozen in to the plasma. The proof of this statement follows by
72 8. Magnetohydrodynamics
deriving the equation of motion for the magnetic eld, and showing the it corresponds to a ow
with velocity u. Before setting = , an equation for the eld follows from Maxwells equation
B/t = curl E with E = u BJ/, which gives
B
t
= curl (u B) +

2
B

, (8.7)
where one uses curl B =
0
J and curl curl B =
2
B for div B = 0. For u = 0, (8.7) reduces to a
diusion equation, with a diusion coecient 1/
0
. Due to this diusion, a magnetic eld line
does not retain its identity: eld lines diuse and ultimately disappear. We are interested in the
case , where the diusion coecient is zero, and the eld lines retain their identity.
A eld line can be dened for any eld. Mathematically, a eld line can be described by the
intersection of two surfaces. In terms of cartesian coordinates, x, y, z, cylindrical polar coordinates,
= (x
2
+ y
2
)
1/2
, , z, and spherical polar coordinates, r, , , these surfaces are solutions of the
equations
dx
B
x
=
dy
B
y
=
dz
B
z
,
d
B

=
d
B

=
dz
B
z
,
dr
B
r
=
rd
B

=
r sin d
B

, (8.8)
respectively. Let us suppose that the two solutions are (t, x) =
0
, and (t, x) =
0
, where

0
,
0
are constants. These are sometimes referred to as Euler potentials. A magnetic eld
line corresponds to the line where the two surfaces intersect for particular values of
0
,
0
. The
magnetic eld can then be written in the form
B = F(, ) grad grad , (8.9)
where F(, ) is a function that depends on the choice of and . One is free to choose the
normalization of , such that one has F(, ) = 1.
The representation (8.9) of the magnetic eld in terms of Euler potentials is convenient for
providing the frozen-in condition. Assuming that the two surface are not changing in shape, but
are moving with a velocity u implies that they satisfy
d
dt
=

t
+u grad = 0,
d
dt
=

t
+u grad = 0, (8.10)
Choosing F(, ) = 1, operating on (8.9) with /t, and using (8.10), gives
B
t
= grad grad[(u grad)] + grad grad[(u grad )]. (8.11)
8.3 Energy density and energy ow 73
The nal steps are left as an exercise, to show that (8.11) implies
B
t
= curl (u B), (8.12)
which is the equation of motion for a eld moving with velocity u. It follows that in (8.7)
implies that the magnetic eld moves at the local uid velocity. This is what is meant by the
statement that the magnetic eld is frozen into the uid.
8.3 Energy density and energy ow
There is no dissipation in ideal MHD, and hence energy and momentum are conserved. The
equation that expresses continuity of energy is of particular interest.
Maxwells equations alone, imply a continuity equation for electromagnetic energy, which is

t
_

0
[E[
2
2
+
[B[
2
2
0
_
+ div
_
E B

0
_
= J E. (8.13)
The terms inside the derivatives on the left hand side are interpreted as the electric energy density,
the magnetic energy density and the electromagnetic energy ux (the Poynting vector), respec-
tively. The right hand term gives the power input per unit volume into the electromagnetic eld.
In ideal MHD one has E+uB = 0, and hence J E = JB u. This allows one to interpret
the right hand side of (8.13) in terms of the rate work is done (per unit volume) by the J B
force in the equation of uid motion (8.2). Assuming = 0, the scalar product of (8.2) with u
can be rearranged, using both the mass continuity equation (8.1) and the adiabatic equation of
state (8.5), so that (8.13) becomes

t
_
u
2
2
+
P
1
+
B
2
2
0
+

0
E
2
2
_
+ div
_
u
2
2
u +
P
1
u +
EB

0
_
= f u. (8.14)
The three terms in the energy density are the kinetic energy density, the thermal energy density,
with P/( 1) the internal energy density, the magnetic energy density and the electric energy
density, respectively. The three terms in the energy ux are the kinetic energy ux, the thermal
energy ux, with P/(1)P the enthalpy, and the the Poynting vector, respectively. The right
hand side is the rate per unit volume that the external force per unit mass, f, does work.
The Poynting vector describes the ux of electromagnetic energy, and in ideal MHD, E =
u B implies
E B

0
=
(u B) B

0
=
B
2

0
u
(u B)B

0
. (8.15)
74 8. Magnetohydrodynamics
Figure 8.1: A front (shaded) region is shown propagating to the right (solid arrow) into a twisted
cylindrical magnetic ux tube; at the front the twist (B

) in the magnetic eld is reduced releasing


magnetic energy. Behind the front the plasma is rotating, corresponding to the unwinding motion
(u

), and there is a Poynting ux (u

B
z
/
0
) directed to the right transporting the energy
released away from the front.
The right hand side includes one term, (B
2
/2
0
)u, that has a simple interpretation: it corresponds
to magnetic energy density being transported at the uid velocity. An example where the nal
term in (8.15) is important is illustrated in Figure 8.1: an unwinding motion can release magnetic
energy that goes into a Poynting vector. In this case the energy ow is along the direction of the
mean background magnetic eld and is orthogonal to the uid velocity.
8.4 Magnetic pressure and magnetic tension
The force (per unit volume) JB in the equation of uid motion (8.2) is the due to the magnetic
stress, sometimes called the Maxwell stress. Using J = curl B/
0
and a vector identity, one has
J B =
(curl B) B

0
= grad
_
B
2
2
0
_
+
(B grad )B

0
. (8.16)
The two terms in (8.16) may be interpreted as a magnetic pressure and a magnetic tension,
respectively.
The magnetic pressure is regarded as isotropic, that is, it acts equally in all directions, just
like the pressure in a gas. The force due to a gradient in the magnetic pressure tends to push
plasma from regions of high magnetic energy density to regions of low magnetic energy density.
The total pressure in a gas is the sum of the magnetic and gas pressures. Consider, for example,
a vertical magnetic ux tube on the surface of the Sun, as illustrated in Figure 8.2. Gravity and
the magnetic tension then both act in the vertical direction. Consider pressure balance in the
8.4 Magnetic pressure and magnetic tension 75
P
m
P
g
grad
P
m
grad
grad P
g
grad
Figure 8.2: A vertical magnetic ux tube above a stellar surface is illustrated schematically. The
magnetic pressure, P
m
, has a gradient directed away from the center of the ux tube, and is
balanced by a gradient in the gas pressure, P
g
, toward the center of the ux tube.
horizontal direction. Pressure balance requires P + B
2
/2
0
= constant in any horizontal plane.
Hence the gas pressure, P = P
out
, outside the ux tube must be greater than the gas pressure,
P = P
in
, inside the ux tube by an amount equal to the magnetic pressure: P
out
P
in
= B
2
/2
0
.
This dierence in gas pressure may be due to a dierence in temperature or to a dierence in
density, or to a combination of the two. For example, sunspots are regions of strong magnetic
eld (B 0.15 T) on the solar surface, and one might speculate that they are darker than their
surroundings because of the gas pressure deciency within the region of strong magnetic eld.
In one sense this must be the correct explanation for why sunspots are dark, but a detailed
justication is much more complicated than this simple suggestion would indicate.
Magnetic tension is closely analogous to the tension in a stretched string. Thus, the eect of
the tension is to tend to reduce the length of the eld line. The tension force may be evaluated
in terms of two unit vectors: b = B/B along the magnetic eld lines, and n pointing towards the
center of curvature of a curved eld line, and dened by writing b grad b = n/R
c
, where R
c
is
the radius of curvature of the eld line. The nal term in (8.16) includes the magnetic tension.
This term may be written in the form
1

0
(B grad)B =
B

0
b grad(Bb) = b b grad
_
B
2
2
0
_
+
B
2

0
n
R
c
. (8.17)
76 8. Magnetohydrodynamics
(a)
y
x
(b)
Figure 8.3: (a) The magnetic eld lines for the eld described in the text are illustrated; the solid
arrow gives the direction of the net tension force. (b) A magnetic conguration containing an
X-type neutral point is illustrated.
The rst term on the right hand side of (8.17) is equal but opposite to the magnetic pressure
gradient in (8.16) along the eld lines. Thus (8.16) and (8.17) imply that the magnetic tension
has a component along the eld line which exactly cancels the pressure gradient along B. The
other term in (8.17) implies a tension force directed toward the center of curvature of the eld
line.
An example of the tension force for a curved eld line is illustrated in Figure 8.3a. The tension
force on the plasma is like that in a slingshot, forcing the plasma in the direction that tends to
reduce the length of the eld lines. The particular eld illustrated in Figure 8.3a corresponds to
B = ay x+b y, where a and b are constants. This corresponds to a current along the z-direction:
J = (/
0
) z. The eld lines are then the surfaces bx +
1
2
ay
2
= constant, z = constant. In this
case the magnetic pressure P
m
= (a
2
y
2
+ b
2
)/2
0
has a gradient only in the y-direction, and the
magnetic tension balances this force in the y-direction and produces a net tension force along the
negative x-axis, given by J B = (ab/
0
) x.
Another example is illustrated in Figure 8.3b. This corresponds to a magnetic eld B =
ay x + bx y, where a and b are constants. The analysis of this example is left as an exercise.
Physically this example may be used to illustrate magnetic reconnection: the central point in
Figure 8.3b is an X-type magnetic neutral point where magnetic eld lines on opposite sides of a
vertical line through the center can drift toward the vertical line, reconnect and drift out as eld
8.5 MHD equilibrium 77
lines on opposite sides of a horizontal line through the center. The net tension force is in the
directions indicated, favoring this reconnection.
8.5 MHD equilibrium
An MHD conguration is said to be in equilibrium if the forces on it are in balance. In most cases
this corresponds to
gradP +J B+f = 0. (8.18)
In modeling a magnetic structure one seeks to nd an appropriate solution of (8.18).
A structure that satises (8.18) is not necessarily stable. As in mechanics, an equilibrium may
be either stable or unstable. There is a large variety of MHD instabilities. Many of these are given
descriptive names, such as the kink instability, the ute instability, the interchange instability, the
sausage instability, the disruptive instability, the sawtooth instability and the ballooning insta-
bility. Such instabilities have been investigated in detail in connection with laboratory plasmas,
where the essential problem is to conne the plasma, and so to prevent instabilities from occurring.
Plasmas tend to be unstable to the simplest and fastest growing instabilities when the magnetic
eld conning the plasma has convex eld lines as viewed from the plasma. Concave eld lines
tend to lead to a more stable conguration. Twisting of eld lines tends to make a conguration
more stable, basically by complicating the topology and so eliminating instabilities that rely on a
simple geometric structure.
The situation with astrophysical plasmas is quite dierent from laboratory plasmas. We see
the astrophysical structures, and so we know that they must be stable in some meaningful sense.
The problem is to explain why the structure is stable, or rather to infer from the incomplete
astrophysical data what the structure might be, granted that it must be relatively stable. Another
type of problem that arises in connection with the stability of astrophysical structures is how to
account for systems that clearly become unstable. A notable case is that of solar magnetic ux
tubes, forming coronal laments or prominences, becoming unstable and leading to the eruption
of the lament or prominence. An eruption involves the lament or prominence lifting o from the
Sun and moving outward through the solar corona and the solar wind. There is a vast literature
on MHD instabilities in general, and of MHD stability in astrophysical plasmas in particular.
However, this topic is not discussed in detail in this lecture course.
78 8. Magnetohydrodynamics
(a) (b)
Figure 8.4: (a) A plasma supported against gravity by a magnetic eld is subject to the Rayleigh
Taylor instability, analogous to a denser liquid supported by a less dense liquid. (b) The instability
develops through ripples at the interface that grow without limit.
Two specic examples of an unstable equilibrium that should be mentioned are the Kelvin
Helmholtz instability and the RayleighTaylor instability. The KelvinHelmholtz instability oc-
curs in a boundary layer between two media in relative motion, and is an important source of
MHD turbulence. The simplest example of a KelvinHelmholtz instability is that of wind over
water generating water waves (which are surface gravity waves). The RayleighTaylor instability
is illustrated in Figure 8.4. The simplest example of such an instability occurs when a less dense
uid (e.g., oil) supports a more dense uid (e.g., water). The nal result of the instability is that
buoyancy causes the less dense uid to ow to above the more dense uid, and the more dense
uid to sink below the less dense uid.
There are several classes of so-called ideal MHD instabilities, which exist in the absence of any
dissipative processes such as electrical and thermal conductivity. Even when an MHD structure
is stable in a nondissipative medium, it may be unstable when dissipative eects are taken into
account. Such instabilities are referred to as resistive instabilities. The important feature that is
permitted when dissipation is included is a relative motion between the uid and the magnetic eld
lines, cf. (8.7). In particular this allows the magnetic topology to change, which is the magnetic
reconnection discussed in connection with Figure 8.3b.
In some astrophysical applications the pressure and gravity forces may be neglected to a rst
approximation, and then (8.18) requires JB = 0. A magnetic structure satisfying this condition
is said to be force free. Thus in a force-free conguration, currents ow along magnetic eld lines.
Somewhere the current much close by owing across eld lines, and hence a force-free structure
must be bounded by a non-force-free region where the currents can close.
8.6 Exercise Set 8 79
8.6 Exercise Set 8
8.1 Show that the magnetic eld in the form (8.9), that is, B = F(, ) grad grad, satises
div B = 0.
8.2 Using the identities
curl (AB) = Adiv B+ (B grad)ABdiv A(A grad)B, (8.19)
and div curl = 0, show that (8.11) implies (8.12).
8.3 The parametric equations for a magnetic eld line are, in cylindrical coordinates,
dx
B
x
=
dy
B
y
=
dz
B
z
. (8.20)
(a) Solve the parametric equation for an idealized magnetic eld corresponding to an X-type
neutral point, see Fig. 8.3b, modeled by B = ay x +bx y, where a and b are constants.
(b) Calculate the current density J = curl B/
0
implied by this eld.
(c) Determine the Maxwell stress for this eld by identifying the magnetic pressure force, grad B
2
/2
0
,
and the tension force, (B grad)B/
0
.
8.4 The magnetic potential for a dipole, m,

M
=
m x
r
3
. (8.21)
(a) Evaluate B = grad
M
in spherical polar coordinates using
B
r
=
d
M
dr
, B

=
1
r
d
M
d
, B

=
1
r sin
d
M
d
. (8.22)
(b) The parametric equations (8.20) for a magnetic eld line are replaced by
dr
B
r
=
rd
B

=
r sin d
B

,
d
B

=
d
B

=
dz
B
z
, (8.23)
in spherical polar coordinates. Write down these equations for a dipolar eld.
(c) Show that a magnetic eld line satises
r = r
0
sin
2
, =
0
, (8.24)
where r
0
and
0
are constants.
80 8. Magnetohydrodynamics
(d) Sketch a eld line satisfying (8.24) indicating the value of r
0
.
8.5 The magnetic potential for a dipole, m, in cylindrical coordinates is

M
=
mz
(
2
+z
2
)
3/2
. (8.25)
(a) Evaluate B = grad
M
in cylindrical coordinates to nd the components
B

=
d
M
d
, B

=
1

d
M
d
, B
z
=
d
M
dz
. (8.26)
(b) Evaluate curl B in cylindrical coordinates, and show that it is identically zero.
(c) Solve the parametric equations
d
B

=
d
B

=
dz
B
z
, (8.27)
to nd the equations for the eld lines in cylindrical coordinates.
Chapter 9
MHD waves
MHD waves are described by solutions of a wave equation derived from the MHD equations.
Waves in magnetized plasmas are well described by MHD theory at frequencies much lower than
the plasma and cyclotron frequencies. In a compressible gas, there are three MHD waves, here
called the Alfven mode and the fast and slow mode.
9.1 Linearized MHD equations
MHD waves correspond to oscillating solutions of the MHD equations. Let the perturbations in
the various variables be denoted by a subscript 1. The only quantities that are assumed to be
nonzero in the absence of the wave are the density, pressure and magnetic eld, and these are
written
=
0
+
1
, P = P
0
+P
1
, B = B
0
+B
1
. (9.1)
All other quantities are assumed to be of rst order in the amplitude of the wave, u
1
, E
1
, J
1
. The
MHD equations are linearized, so that each term is of rst order in a perturbed quantity. The
equations are then expanded in plane waves. Applying this procedure to the equation of mass
continuity (8.1) gives
i
1
+i
0
k u
1
= 0. (9.2)
Applying the procedure to the equation of uid motion (8.2), with the charge density and the
external force assumed to be zero, gives
i
0
u
1
= J
1
B
0
. (9.3)
82 9. MHD waves
The electric eld in ideal is given by E +u B = 0, which linearizes to
E
1
= u
1
B
0
. (9.4)
Only the rst two of Maxwells equations are needed, and with the displacement current neglected,
these become
k E
1
= B
1
, ik B
1
=
0
J
1
. (9.5)
We consider only incompressible and adiabatic equations of state, given by the rst and last of
(8.5), respectively, and these linearize to
k u
1
= 0, P
1
= c
2
s

1
(9.6)
with c
2
s
= P
0
/
0
the adiabatic sound speed.
Our objective is to reduce these equations to a single MHD wave equation, and solve it for the
wave properties. To derive the wave equation we need to choose one variable and to express all the
other variables in terms of the chosen variable. One possible choice is u
1
, and this is equivalent
to choosing the uid displacement . This corresponds to
u
1
= i, (9.7)
where the subscript 1 on is omitted. It is convenient to write
B
0
= B
0
b, v
2
A
=
B
2
0

0
, (9.8)
where v
A
is the Alfven speed.
9.2 Alfven mode
Alfven waves are analogous to waves on a stretched string, with the magnetic tension replacing
the tension in the string. These waves involve no compression of the uid, and so that exist in an
incompressible uid. It is relatively simple to derive their properties for an incompressible gas.
In the incompressible case (9.7) implies k u
1
= 0, and then (9.2) implies
1
= 0, Eliminating
J
1
between (9.3) and the second of (9.5) gives i
0
u
1
= i(k B
1
) B
0

0
, and writing B
1
in
terms of E
1
using the rst of (9.5), and thence in terms of u
1
using (9.4) gives an equation that
9.3 Magnetoacoustic waves 83
involves only u
1
. In this incompressible case, if one writes u = i
A
, and k = k, (9.2) and
(9.3) imply

A
= 0, b
A
= 0,
A
b, (9.9)
where the third relation follows from the rst two. The resulting dispersion relation for Alfven
waves is

2
= k
2
v
2
A
cos
2
, cos b. (9.10)
One also has
u
1
= i
A
E
1
= iB
0

A
b,
B
1
= ikB
0
cos
A
, J
1
=
k
2
B
0
cos

0

A
. (9.11)
In an Alfven wave, the uid velocity, the electric eld, the magnetic uctuation and the
current are all in the plane orthogonal to the background magnetic eld. When uid is displaced
perpendicular to B
0
, the magnetic eld is displaced with the uid, the eld line becomes locally
curved, and this leads to a tension tending to straighten the eld line by pulling the uid element
back toward the unperturbed conguration. The inertia of the plasma causes it to overshoot,
setting up an oscillatory motion. The speed of wave propagation may be understood by analogy
with the speed of energy propagation for a stretched string: the wave velocity on a stretched string
is [(tension)/(mass per unit length)]
1/2
, which may be rewritten as [(tension per unit area)/(mass
per unit volume)]
1/2
. The magnetic tension per unit area is B
2
0
/
0
and the mass per unit volume
is
0
, implying that the wave speed is v
A
.
9.3 Magnetoacoustic waves
The properties of magnetoacoustic waves may be derived from the set of equations used to treat
Alfven waves, but with the incompressible condition relaxed. This is appropriate provided that
the magnetic pressure is much greater than the gas pressure, so that the perturbations in and
P are unimportant. On writing u = i, one infers that all solutions must have b = 0.
According to (9.9), Alfven waves have
A
b. There is another possible solution that is
orthogonal to this, specically, one with =
M
,

M
bb. (9.12)
84 9. MHD waves
The solution corresponding to (9.12) describes magnetoacoustic waves, which satisfy the dispersion
relation

2
= k
2
v
2
A
. (9.13)
Unlike Alfven waves, these waves involve a magnetic compression; that is, B
1
= i (
M
B
0
)
has a component along B
0
.
9.4 Fast and slow modes
The linearized MHD equations in a compressible plasma lead to the MHD wave equation, which
may written in the matrix form
= 0, (9.14)
where is the uid displacement vector, and with the matrix of coecients
=
2
1 k
2
c
2
s
k
2
v
2
A
[ cos (b +b) + cos
2
1]. (9.15)
We are free to choose the coordinate axes such that
= (sin , 0, cos ), b = (0, 0, 1), (9.16)
and then the matrix form for is
=
_
_
_
_
_
_
_

2
k
2
v
2
A
k
2
c
2
s
sin
2
0 k
2
c
2
s
sin cos
0
2
k
2
v
2
A
cos
2
0
k
2
c
2
s
sin cos 0
2
k
2
c
2
s
cos
2

_
_
_
_
_
_
_
. (9.17)
The dispersion relation for MHD waves may be found by setting its determinant of equal to
zero. This gives
(
2
k
2
v
2
A
cos
2
)[
4

2
k
2
(v
2
A
+c
2
s
) + k
4
v
2
A
c
2
s
cos
2
] = 0. (9.18)
One solution of (9.18) corresponds to
2
k
2
v
2
A
cos
2
= 0, and this corresponds to Alfven waves.
The other factor describes the fast and slow modes.
It is convenient to solve for the square of the phase speed, v
2

=
2
/k
2
. The solutions are
v
2

= v
2

, v
2

=
1
2
(v
2
A
+c
2
s
)
1
2
[(v
2
A
+c
2
s
)
2
4v
2
A
c
2
s
cos
2
]
1/2
. (9.19)
9.4 Fast and slow modes 85
These correspond to the fast (+) and slow () magnetoacoustic modes.
The fast and slow modes may be interpreted in terms of gas sound waves modied by the
contribution from the magnetic pressure, and magnetoacoustic waves modied by the contribution
from the gas pressure. The fast mode is gas sound-like for c
2
s
v
2
A
and magnetoacoustic-like for
c
2
s
v
2
A
. Conversely, the slow mode for propagation nearly along the eld lines is magneto-
acoustic-like for c
2
s
v
2
A
and gas-sound-like for c
2
s
v
2
A
, and is quite strongly modied from these
properties at large angles of propagation. In the low- limit, (9.19) may be approximated by
v
2
+
v
2
A
+c
2
s
sin
2
, v
2

c
2
s
cos
2
. (9.20)
The fast mode is a slightly modied magnetoacoustic mode and the slow mode reduces to a sound
wave for propagation along the magnetic eld lines, but is increasingly modied by the magnetic
eld as the direction of propagation approaches perpendicular to the eld lines.
The direction of the uid displacement may be found from the wave equation (9.14). For
Alfven waves one nds
A
= (0, 1, 0), as given by (9.9). The direction of the uid displacement in
the magnetoacoustic waves is given by

= (sin

, 0, cos

), tan

=
c
2
s
sin cos
v
2

v
2
A
c
2
s
sin
2

. (9.21)
Using these properties one may nd the other elds in the waves from the linearized MHD equa-
tions.
86 9. MHD waves
9.5 Exercise Set 9
9.1 In very low density plasmas the displacement current cannot be neglected when treating MHD
waves.
(a) Show that when the displacement current is included (9.15) is replaced by
=
2
1 +
v
2
A
c
2
(1 bb) k
2
c
2
s
k
2
v
2
A
[ cos (b +b) + cos
2
1]. (9.22)
(b) Hence show that the dispersion relation for Alfven waves, =
A
(k), becomes

A
(k) =
kv
A
[ cos [
(1 +v
2
A
/c
2
)
1/2
. (9.23)
(c) As the plasma density approaches zero, n 0, the Alfven speed approaches innity, v
A
.
Show that the dispersion relation for Alfven mode approaches = kc[ cos [ in this limit, and nd
the corresponding limiting value of the dispersion relation for the fast mode (assuming c
s
c).
(d) One expects the wave properties to approach those of electromagnetic waves in vacuo in this
limit, but the dispersion relation for Alfven mode approaches = kc[ cos [ and not = kc. Can
you suggest how this dilemma might be resolved?
9.2 Show that the group velocity for Alfven waves is along the magnetic eld irrespective of the
angle .
Hint: Write the dispersion equation =
A
(k) with
A
(k) = [k
z
[v
A
and evaluate the components

A
(k)/k
x
,
A
(k)/k
y
,
A
(k)/k
z
, and interpret the result.
Chapter 10
Collisions (Coulomb interactions)
Wave damp in any medium due to dissipative eects. Collisions between electrons and ions leads
to dissipation, causing waves to damp. Collisions in a plasma are qualitatively dierent from
collisions in an un-ionized gas. The simplest model for collisions between atoms (or molecules)
involves approximating each atom by a hard sphere, and considering how these bounce o each
other. Such a model is completely invalid for collisions in a plasma. The simplest useful model
involves assuming that collisions are due to Coulomb interactions, with the Coulomb eld modied
by Debye screening.
10.1 Coulomb interactions
In a plasma, charged particles interact via their Coulomb elds, provided their separation is less
than about a Debye length. In the case of a Coulomb interaction between an electron and an ion,
the mass of the ion is so much greater than that of the electron, that the ratio can be assumed
innite to a rst approximation. The motion of the electron is then that of a particle in an inverse-
square-law eld. This problem is familiar from the study of the motions of the planets about the
Sun. The orbits are conic sections with the Sun at one of the foci. Conic sections can be ellipses or
hyperboli (with a circle and a parabola being special cases). The planets move in elliptical orbits,
and objects that come from outside the solar system move in hyperbolic orbits. The motion of an
electron about an ion in a Coulomb interaction can be approximated by an hyperbola, provided
that the distance of closest approach in less than the Debye length.
Consider a Coulomb interaction between an electron with charge e and an ion with charge
88 10. Collisions (Coulomb interactions)
Z
i
e. In the approximation in which the electron-ion mass ratio is assumed innite, the ion is
eectively a xed scattering center. An electron with initial v is assumed to pass the scattering
center with impact parameter b, resulting in a nal velocity v

. The energy of the electron in the


initial and nal states is the same, implying v

= v. The scattering angle, , is dened by


v v

= v
2
cos . (10.1)
The relation between and the impact parameter for a Coulomb interaction is
tan
1
2
=
b
0
b
, b
0
=
Z
i
r
0
(v/c)
2
, (10.2)
where r
0
= e
2
/4
0
m
e
c
2
is the classical radius of the electron. The parameter b
0
corresponds to
the impact parameter that leads to a deection through = /2.
The dierential cross section for scattering into an element d
2
of solid angle is d/d
2
=
b db/d cos , which corresponds to the Rutherford cross section,
d
d
2

=
b
2
0
4 sin
4 1
2

. (10.3)
For large scattering angles, /2, this cross section is approximately b
2
0
. For small , the
cross section increases 1/
4
as decreases. This implies a cross section b
4
/b
2
0
for b b
0
.
If follows that the most frequent scattering are those with the largest b. Due to Debye screening
the largest possible impact parameter is b
D
. This corresponds to small-angle scattering,
2b
0
/b 2b
0
/
D
.
10.2 Straight-line approximation
Consider the frequency of large-angle collisions between electrons and ions. An electron with speed
v has encounters with b b
0
at a rate n
i
v, where n
i
is the number density of ions, and where
the cross section can be estimated as b
2
0
for such interactions. Each such interaction causes
deection through a large angle, and this rate may be interpreted as the collision frequency for
large-angle scattering. Using (10.2), this gives a collision frequency Z
2
i
n
i
r
2
0
c
4
/v
3
=
4
p
/n
e
v
3
,
where the nal formula applies for Z
i
= 1, n
i
= n
e
. However, scattering by the large number
of distant encounters, with b
D
, is more ecient than such relatively infrequent large-angle
scattering. As a result, the eective collision frequency is larger than this estimate for large-angle
scattering suggests.
10.3 Collision frequency 89
P
P
O
b
b vt +
0
Figure 10.1: The orbit of an electron in the straight-line approximation: the ion is located at the
origin O, and the position of the electron is x = b +vt, where b = [b[ is the impact parameter.
A major simplication that occurs in the case of distant encounters is that each encounter
causes only a tiny change in , so that the motion of the electron may be treated as constant
linear motion to s rst approximation. This is called the straight-line approximation, as illustrated
in Figure 10.1. The many small changes in , due to the large number of distance encounters,
may be treated using a diusion approximation: the direction of propagation changes slowly and
randomly due to the many small-angle scatterings.
10.3 Collision frequency
Coulomb interactions play the role of collisions in a plasma. Scattering of electrons by ions tends
to make the electrons isotropic. Suppose the electrons are streaming relative to the ions. Then
the scattering of the electrons out of the streaming direction tends to reduce the relative velocity
between the electrons and the ions. Such a relative ow corresponds to an electric current, and
the reduction in the ow reduces the current, corresponding to an electrical resistivity. Scattering
also tends to transfer energy between particles. In plasmas the temperature of the electrons and
of the ions is often not the same, due to various heating mechanisms usually favoring one species
over another. Coulomb interactions tend to reduce the temperature dierence, by transferring
90 10. Collisions (Coulomb interactions)
b
b
+
d
b
v
Figure 10.2: An electron has distant encounters with ions with impact parameter between b and
b +db at a rate n
i
v2b db per unit time
energy on average from the hotter to the cooler component. A fast particle moving through a
plasma tends to lose energy to the thermal particles as a result of Coulomb interactions.
A detailed theory of Coulomb interactions involves the use of kinetic theory, with the electron
distribution evolving due to diusion in angle and in speed. The rate at which the electrons
diuse in angle, due to distant encounters, is described by a collision frequency that is estimated
as follows. The change in angle in a distant encounter is 2b
0
/b 1, where (10.2) is used.
As illustrated in Figure ??, the rate of such encounters in the range b to b +db is estimated from
n
i
v2bdb. The rate of diusion in angle is n
i
v()
2
2bdb. On integrating over impact parameters,
this rate involves a logarithmic integral,
_
db/b. Cutting this integral o at b < b
0
and at b >
D
,
it gives the Coulomb logarithm:
_

D
b
0
db
b
= ln
c
, (10.4)
with
c
may be approximated by the the Debye number, (2.13). Apart from a factor of order
unity, this calculation gives the collision frequency

c
=

i
Z
2
i
n
i
e
4
4
2
0
m
2
e
v
3
ln
c
, (10.5)
10.4 Electrical conductivity and runaway electrons 91
where a sum over ionic species is included. For a hydrogen plasma, Z
i
= 1, n
i
= n
e
, (10.5) gives

c
=
p
_
V
e
v
_
3
ln
c

c
. (10.6)
A more detailed calculation shows that this collision frequency applies to both diusion in angle
and diusion in v.
The collision frequency corresponding to distant encounters is larger than that for large-angle
scattering by a factor ln
c
. It follows that the frequent encounters with b
D
, each leading
to a small-angle scattering, have a combined eect that is larger than the relatively infrequent
large-angle scattering events.
Note that for both the large-angle and small-angle scattering, the collision frequency decreases
with increasing speed, v, of the electron. The dependence
c
1/v
3
implies that very fast electrons
experience very few collisions.
10.4 Electrical conductivity and runaway electrons
Collisions lead to dissipation, and various kinds of dissipation can be described by dierent trans-
port coecients. Here only one is considered: electrical resistivity. Other transport coecients
include thermal conductivity, viscosity, diusivity, and so on. All these can be calculated for
plasmas in terms of Coulomb interactions.
An electric eld in a conductor drives a current, due to electrons moving relative to ions. The
velocity of an electron increases as it is accelerated by the electric eld, until it has a collision and
its velocity is randomized. The randomizing eect of collisions can be described by a drag term
in the equation of motion for the electrons:
m
e
dv
dt
= eE
c
m
e
v. (10.7)
The acceleration and the drag are in balance when the electrons drift with a drift speed v = v
d
,
v
d
= eE/
c
m
e
. (10.8)
The current density is then J = en
e
v
d
, which is proportional to E. The constant of proportion-
ality is the electrical conductivity,
e
, given by

e
= e
2
n
e
/
c
m
e
=
0

2
p
/
e
. (10.9)
92 10. Collisions (Coulomb interactions)
A surprising result emerges when one considers electrons with dierent speed, and takes the
dependence 1/v
3
of the collision frequency on speed. There are always some electrons with high
enough speed that collisions are ineective in preventing them being accelerated. These electrons
continue to be accelerated and the faster they become, the fewer collisions there are to slow them
down. This leads to an important phenomenon is plasma: runaway electrons. These are electrons
that are accelerated to high velocities by any electric eld in the plasma. Such runaways can be
a serious problem in plasma machines.
10.5 Damping of waves
A wave oscillates as exp[i(t k x)], and its damping may be described by giving or k an
imaginary part. Whether waves damp in time or space depends on the boundary conditions. If
the waves are uniformly excited everywhere at an initial time, they damp in time, and if the waves
are excited by a point source emitting at a constant rate, they damp in space away from the point
source. Consider purely temporal damping. With Im the imaginary part of , the amplitude
varies secularly with time as exp(Im t). The energy in the waves is proportional to the square
of the amplitude, and if it damps as exp(t), where is the absorption coecient, then one has
= 2Im. On including spatial damping, the absorption coecient is related to the imaginary
parts by
= 2(Im v
g
Imk), (10.10)
where v
g
is the group velocity.
Collisions lead to dissipation, and dissipation of waves is described in terms of damping. For
Langmuir waves and transverse waves in a plasma, the following argument allows one to determine
for given collision frequency,
c
. The physical argument concerns the ratio of the energy W
K
/W
T
.
The energy W
K
is in forced motion of the particles, associated with their response to the wave.
In Langmuir waves and transverse waves only the forced motion of the electrons is important.
Consider the eect of one of the electrons having a collision. Its motion is then randomized, and
hence the energy that its forced motion contributed to the wave is lost. It follows that the waves
lose energy at the rate
c
W
K
. By assuming that the waves damp with an absorption coecient
(omitting the subscript M for the present) implies that the total energy in the waves decreases at
10.5 Damping of waves 93
the rate W
T
. Equating these two rates of energy loss gives
=
c
_
W
K
W
T
_
, (10.11)
where W
K
/W
T
is the ratio of kinetic to total energy in the waves. For Langmuir waves this ratio
is close to 1/2, and hence the collisional absorption coecient is
c
/2. For transverse waves the
ratio is given by (4.27), implying an absorption coecient
c

2
p
/2
2
.
94 10. Collisions (Coulomb interactions)
10.6 Exercise Set 10
10.1 Find the absorption coecient for collisional damping for Langmuir waves and transverse
wave, assuming that each collision thermalizes the fraction W
K
/W
T
of the wave energy, with this
fraction given by (4.3) or (4.4).
10.2 Show that the Coulomb logarithm calculated using (10.4) diers by only a small additive
constant from the logarithm of the Debye number (2.13). This number is unimportant for typical
values, ln
c
1020.
Chapter 11
Landau damping
There is also a form of collisionless damping in a plasma. This is referred to as Landau damp-
ing.
1
This form of damping is due to resonant interactions between waves and particles. Landau
damping can be negative, correspond to wave growth rather than damping.
11.1 Landau damping
Collisionless damping, called Landau damping, is due to wave-particle resonance. The resonance
condition, given by (3.8), is k v = 0. This condition can be satised only if the phase speed,
v

= /k, satises v

v < c. The phase speed of transverse waves exceeds the speed of light,
and because no resonance is possible, there is no Landau damping of transverse waves. Landau
damping of Langmuir waves plays an important role in plasma physics.
One can understand how Landau damping occurs by considering the eect of the resonant
interaction on the electrons. When the resonance condition is satised, we can imagine how an
observer moving with the phase velocity of the wave sees the interaction. Such an observer sees
the wave as stationary, oscillating in space but not in time. A resonant electron electron is at rest
relative to our moving observer. The electric eld accelerates a resonant electron causing it to
gain energy. Where does this energy come from? Energy is conserved, so it must come from the
wave. Hence the wave damps.
This explanation is qualitatively correct, but it too oversimplied to understand the physics
properly. This is because the probability of an electron being exactly in resonance is negligibly
1
Landau pointed out the existence of this eect in a paper in 1946.
96 11. Landau damping
small, and one needs to consider electrons in a small but nonzero range of velocities about the
resonant velocity. Some of these electrons are moving more slowly than the wave, and some are
moving faster than the wave. Analyzing the eect of the wave on these electrons, one nds that on
average the eect is to draw the electrons into resonance. This means that the slower electrons gain
some energy from the wave, and the faster electrons lose some energy to the wave. Landau damping
occurs if there are more slower electrons than faster electrons. For a Maxwellian distribution,
f(v) exp(v
2
/2V
2
e
), where V
e
= (k
B
T
e
/m
e
)
1/2
is the thermal speed of the electrons, one has
df(v)/dv < 0, implying that there are more slower than faster electrons. Hence damping occurs
for a Maxwellian distribution. The absorption coecient for Landau damping by a Maxwellian
distribution is proportional to exp(v
2

/2V
2
e
), where v

is the phase speed of the wave.


11.2 Dispersion and dissipation
A full understanding of Landau damping involves a subtle part of physics involving dispersion
theory, causality and the second law of thermodynamics. Landau imposed the cause condition
explicitly: the response of the plasma cannot occur before the disturbance that causes it. This
relation implies that any causal function, describing such a response, has a specic integral relation
between its real and imaginary parts as a function of frequency. Separating the response function
into real and imaginary parts by writing
K
L
(, k) = Re K
L
(, k) + i ImK
L
(, k), (11.1)
the so-called Kramers-Kronig relations imply
Re K
L
(, k) 1 =
1

T
_
d

ImK
L
(

, k), (11.2)
ImK
L
(, k) =
1

T
_
d

[Re K
L
(

, k) 1], (11.3)
where T denotes the Cauchy principal value of the integral. The absorption coecient for Lang-
muir waves can be written

L
(k) = 2
L
(k)R
L
(k) ImK
L
(
L
(k), k), (11.4)
where R
L
(k) is the ratio of electric to total energy in the waves. It is evident from this line of
reasoning that Landau damping is somehow associated with causality.
11.3 Negative Landau damping 97
0 10
F(v)
v
Figure 11.1: A bump-in-tail distribution F(v). Waves with phase speed equal to the particle speed
either grow or damp due to the eect of the wave-particle resonance; growth occurs in the range
dF(v)/dv > 0.
A related approach is based on the second law of thermodynamics. As with causality, this
law builds the sense of time into a physical description. Langmuir waves are emitted by the
plasma electrons through a resonant interaction satisfying k v = 0. This is sometimes called
the Cerenkov condition, and the emission process is referred to as Cerenkov emission. Without
performing any calculations, the second law requires that there be a corresponding absorption
process. If this were not the case, the level of the Langmuir waves would build up without limit.
In thermal equilibrium the level of the waves must be the thermal level, which corresponds to
W(k) = T
e
. The absorption and emission must be in balance (for thermal electrons) to produce
this level. The absorption process is Landau damping, which may be interpreted as the absorption
process corresponding to Cerenkov emission.
11.3 Negative Landau damping
Landau damping is negative for a distribution in which the distribution function is increasing
as a function of velocity v = v

, where v

is along the direction k. When this condition is


satised, energy ows from the electrons to the waves, causing the waves to grow. An example
98 11. Landau damping
Figure 11.2: A dynamic spectrum of solar radio bursts following a solar are. Frequency is plotted
from high to low, to show a disturbance moving outward through the solar corona.
of a distribution function for which this occurs is shown in Figure 11.2. This bump-in-tail
distribution consists of a background Maxwellian distribution and a beam of faster electrons
propagating through the plasma. The growth of the waves is exponential, and it continues until
a substantial fraction of the energy initially in the electron beam is transferred to the Langmuir
waves.
Runaway electrons tend to form an electron beam in a plasma, and cause Langmuir waves to
grow.
Growth of Langmuir waves plays a central role in plasma emission in solar radio bursts.
Following a solar are, radio observations of the Sun show bursts that drift rapidly from high
to low frequency. These burst correspond to emission from the solar corona at the local plasma
frequency and its second harmonic. The drift from high to low frequency is due to the electron
beam, generated in the solar are, propagating outward, from higher to lower densities. These
electron beams propagate through the interplanetary plasma, and can be observed in situ by
spacecraft. The Langmuir waves generated by the electron beam cannot escape from the plasma.
The observed emission is due to nonlinear processes in the plasma, partially converting the energy
in Langmuir waves into transverse waves, at
p
and 2
p
. Runaway electrons tend to form an
electron beam in a plasma, and cause Langmuir waves to grow. The Langmuir waves can scatter
the electrons, preventing their further acceleration, producing a form of anomalous conductivity.
Growth of Langmuir waves due to a beam is only one example of a plasma instability. A related
instability occurs due to a current owing in a plasma. A current requires the electrons moving
11.3 Negative Landau damping 99
relative to the ions. If this relative motion exceeds the ion sound speed, Landau damping becomes
negative for ion sound waves, and these waves grow. The ion sound waves scatter the electrons,
impeding the current ow. This implies a collisionless form of electric resistivity, referred to as
anomalous resistivity. The value of the anomalous conductivity is often represented by the form
(10.9) with ,=
e
replaced by
p
, with a number of order 0.1.
One way of heating a fusion plasma involves driving a beam of particles into the plasma, relying
on a beam instability to transfer the energy in the beam to wave through a resonant instability,
with the waves damping and heating the plasma. A beam of neutral atoms is injected, and when
the atoms become ionized in the plasma, this becomes an ion beam capable of generating waves
through negative Landau damping.
100 11. Landau damping
11.4 Exercise Set 11
11.1 Collisions are included in the longitudinal response function in the cold plasma limit by
writing
K
L
() = 1

2
p
( +i
c
)
. (11.5)
Identify the imaginary part and used (11.4) to calculate the absorption coecient.
11.4 Exercise Set 11 101
Ftensor
102 11. Landau damping
Chapter 12
Kinetic theory
Modern plasma physics is based on a statistical theory called the kinetic theory of plasmas. In
this lecture the statistical ideas that underlie this theory are introduced.
12.1 Phase space and distribution functions
Consider the air in this room. It consists of a large number of molecules, mainly N
2
and O
2
,
which for simplicity we regard as identical structureless particles of mass m. Any given particle
has a velocity v and momentum p = mv. The particles propagate in approximately straight lines
between collisions with each other and with the walls, when their momentum changes abruptly. In
a statistical description we imagine a 6-dimensional phase space consisting of the three components
of position, x, and the three components of momentum, p. An individual particle at any one
time is described by a point in this phase space. As time evolves, this point moves through
the phase space, with p constant except when it has a collision with another particle or a
wall and its momentum changes rapidly to a new value. Mostly we are not interested in the
details of the individual particles. Assuming a very large number of particles, we replace the
distribution of points moving through the phase space by a smoothed distribution. Let f(p, x, t)
be the distribution function such that the number of representative points in an innitesimal
volume of the phase space is d
3
xd
3
p is f(p, x, t) d
3
xd
3
p. (The notation d
3
x means dx dy dz and
d
3
p means dp
x
dp
y
dp
z
in cartesian coordinates.) Suppose we consider some volume, say a cubic
meter, of the air in this room. The distribution function for the particles in this volume is very
well approximated by a Maxwellian distribution at the temperature of the room. Note that this
104 12. Kinetic theory
distribution does not change signicantly as a function of position of time. Although the individual
particles move through the volume, the distribution function does not change. The dependence of
the distribution function on x, t does not describe the motion of individual particles. What this
dependence describes can be understood by thinking of a snapshot of the particles in a volume
centered on one point at one time, and comparing it with a snapshot of the particles in a similar
volume centered on a dierent point at a dierent time. If these snapshots are identical, then the
distribution function is independent of position and time. The snapshots of the particles in small
volumes of the air in this room are identical on short time scales, but change slowly through the
day, due to slow changes in the temperature, and due to drafts and other driven movements of
the air.
The distribution function is normalized to the number of particles, N, in the volume, V :
N =
_
d
3
xd
3
pf(p, t, x). (12.1)
The number density, n = N/V , of particles of a given species is determined by
n(t, x) =
_
d
3
pf(p, t, x). (12.2)
We are concerned with the current and charge densities due to charged particles. For a species
with charge q, these are given by
(t, x) = q
_
d
3
pf(p, t, x), J(t, x) = q
_
d
3
pv f(p, t, x), (12.3)
respectively.
The mean distribution function, f
0
(p) say, is found by averaging over uctuations in time and
space. Let the mean number density be n
0
= N/V . The normalization condition is
n
0
=
_
d
3
pf
0
(p). (12.4)
A Maxwellian distribution corresponds to f
0
(p) exp(p
2
/2mk
B
T), where T is the temperature.
Carrying out the integral in (12.4) determines the constant of proportionality, giving
f
0
(p) =
n
0
(2mk
B
T)
3/2
exp
_

p
2
2mk
B
T
_
. (12.5)
12.2 Boltzmann and Vlasov equations 105
12.2 Boltzmann and Vlasov equations
Slow changes of the distribution function are described by Boltzmanns equation. (The Boltzmann
equation is derived from Liouvilles theorem, which is that the distribution function is a constant
along the trajectory in phase space; here we take the Boltzmanns equation as given.) In the
absence of collisions, the (collisionless) Boltzmanns equation is
_

t
+v

x
+F(t, x)

p
_
f(p, t, x) = 0, (12.6)
where F(t, x) is the force acting on a particle, such that its equation of motion is dp/dt =
F(t, x). The sum of dierential operators on the left hand side of (12.6) corresponds to the total
time derivative expressed as as a convective derivative along the trajectory. (The Lorentz force,
qv B(t, x), depends on p, and it is possible to show that (12.6) is also valid in this case.) There
is one Boltzmann equation for each dierent species of particle. For a collisionless system, this
implies that the total time derivative of the distribution function is zero.
An important point in understanding how (12.6) is re-interpreted in plasma kinetic theory is
that the force in (12.6) is an external force. Then (12.6) is a linear partial dierential equation
for f(p, t, x). In plasma kinetic theory, the force is identied as
F(t, x) = q[E(t, x) +v B(t, x)], (12.7)
with E(t, x), B(t, x) interpreted as the actual elds in the plasma. These elds depend on the
charge and current densities, which depend on the distribution function. These are called the self-
consistent elds. The simplest example of a change from an external eld to a self-consistent eld
is in Debye shielding. In this case the external eld is the Coulomb eld due to the charge, and
the charge density induced in the plasma leads to another contribution to the eld, which is found
by solving Gauss equation with this additional charge density as the source. The self-consistent
eld is the resulting eld that describes a charge that is screened at distance much greater than
the Debye length.
The basic equations of the kinetic theory of plasmas are called the Vlasov equations. These
consist of the following:
1. A set of equations which is formally identical to Boltzmanns equation for each species of
particles, as in (12.6):
_

t
+v

x
+q

[E(t, x) +v B(t, x)]



p
_
f

(p, t, x) = 0, (12.8)
106 12. Kinetic theory
where f

(p, t, x) is the single particle distribution function for species .


2. A pair of equations giving the charge and current density in terms of the single particle
distribution functions, as in (12.3):
(t, x) =

_
d
3
pf

(p, t, x), J(t, x) =

_
d
3
pv f

(p, t, x), (12.9)


where the sum is over all species of particle.
3. Maxwells equations with the source terms identied with those in (12.9).
Equations (12.8) are a set of non-linear coupled integro-dierential equations for the distri-
bution functions f

. One can solve them only by making simplifying approximations. The ap-
proximation made here is to expand in powers of the uctuating elds. To zeroth order in this
expansion, the distribution functions are the mean distributions. The terms of rst order in the
uctuating elds are referred to as the linearized terms.
12.3 Linearized Vlasov equation
The linear response is found by expanding the distribution functions in powers of the electromag-
netic disturbance,
f

(p, t, x) = f
0

(p) + f
(1)

(p, t, x) + , (12.10)
where f
0

(p) is the mean distribution function, and where + denotes higher order terms that
are ignored.
The linearized Boltzmann equation follows from (12.8):
_

t
+v

x
_
f
1

(p, t, x) + q

[E(t, x) +v B(t, x)]



p
f
0

(p) = 0. (12.11)
One expands (12.11) in plane waves. Formally, the expansion in plane waves corresponds to
Fourier transforming, and for some purposes it is important to appeal to the known properties
of Fourier transforms, which are summarized in Appendix B. For most purposes it suces to
write a uctuating quantity, f(t, x), as

f(, k) exp[i(t k x)], and to interpret

f(, k) as the
12.4 Tensor notation 107
uctuating Fourier components. The Fourier transform of Maxwells equations follow from (4.2)
simply by adding the arguments (, k). The Fourier transform of the force (12.7) becomes

F(, k) = q[

E(, k) +v

B(, k)]. (12.12)
Using the rst of Maxwells equations (4.2), (12.12) becomes

F(, k) =
q

[( k v)

E(, k) + kv

E(, k)]. (12.13)
Then (12.11) gives
i( k v)f
(1)

(p, , k) +
q

[( k v)

E(, k) +kv

E(, k)]

p
f
0

(p) = 0. (12.14)
The solution of (12.14) allows one to calculated the uctuating charge and current densites using
the Fourier transforms of (12.9):
(, k) =

_
d
3
pf
(1)

(p, , k),

J(, k) =

_
d
3
pv f
(1)

(p, , k). (12.15)


The charge and current densities are related by the continuity equation (4.3). It suces to calculate
the current

J(, k).
12.4 Tensor notation
The induced current,

J(, k), is a linear function of the electric eld,

E(, k), but in general it is
not parallel to the electric eld. To describe the relation it is useful to introduce tensor notation.
The notation is summarized in Appendix A. Omitting the tildes, the relevant linear relation is
J
i
(, k) =
ij
(, k)E
j
(, k), (12.16)
where
ij
(, k) is the conductivity tensor. One can write (12.16) in matrix form, with the three
components of J
i
(, k), with i = x, y, z, and of E
j
(, k), with j = x, y, z, are written as a column
matrices, with
ij
(, k) written as a square matrix with component i, j = x.y.z.
In place of the conductivity tensor
ij
(, k), it is often more convenient to use the equivalent
dielectric tensor, K
ij
(, k) to describe the response. The equivalent dielectric tensor is dened to
be dimensionless and to reduce to the unit tensor in vacuo. One denes K
ij
(, k) by writing
K
ij
(, k) =
ij
+
i

ij
(, k), (12.17)
108 12. Kinetic theory
where
ij
is the unit tensor (or unit matrix in matrix notation). The equivalent permittivity tensor

ij
(, k) =
0
K
ij
(, k) diers from the dielectric tensor only in its dimensions.
On may write down an expression for the conductivity tensor by inserting (12.14) into (12.15).
It is useful to rst rewrite (12.14) in tensor notation:
i( k v)f
(1)

(p, , k) =
q

[( k v)
sj
+k
s
v
j
]E
j
(, k)

p
s
f
0

(p). (12.18)
The resulting expression for the equivalent dielectric tensor is
K
ij
(, k) =
ij
+

q
2

2
_
d
3
p
v
i
k v
[( k v)
sj
+k
s
v
j
]

p
s
f
0

(p). (12.19)
The result (12.19) is the general form for the response tensor for an arbitrary plasma in the absence
of a static magnetic eld.
12.5 Isotropic plasma
For an isotropic medium f
0

(p) depends only on the magnitude p = [p[ and not on its direction.
In this case K
ij
(, k) must be of the form
K
ij
(, k) = K
L
(, k)
i

j
+K
T
(, k) (
ij

j
), (12.20)
The derivation of (12.20) involves a subtle but powerful argument in tensor analysis. The argument
is as follows. The tensor can be expressed as a sum of scalars times basic tensors. There are at
most nine scalars (corresponding to the nine components of a 33 matrix), and the tensor satised
Onsager relations between its o-diagonal components such that there are only six independent
components. In an isotropic medium there is no natural direction. If K
ij
did not depend on k
there would be no direction in the problem. and the only possibility is that K
ij
is a scalar times
the unit tensor
ij
. In this case there is only one independent component. The dependence on
k introduces a vector into the problem, and one can form another tensor,
i

j
, that satises the
Onsager relations. The components along
ij
and
i

j
can be dierent, and this is what is assumed
in (12.20). (There is another tensor,
ijl

l
that satises the Onsager relations, and this allows a
rotatory component, which is zero in a plasma, but nonzero in an optically active medium such
as a solution of dextrose.)
The longitudinal and transverse parts are obtained from (12.19) using
K
L
(, k) =
i

j
K
ij
(, k), K
T
(, k) =
1
2
(
ij

j
)K
ij
(, k). (12.21)
12.6 Thermal plasma 109
Assuming an isotropic distribution implies that f
0

(p) depends only on the magnitude of p.


Then in (12.14) one has

p
f
0

(p) =
v
v

p
f
0

(p), f
(1)

(p, , k) =
iq

v

E(, k)
v( k v)

p
f
0

(p), (12.22)
The induced current (12.15) becomes

J(, k) =

iq
2

_
d
3
p
v v

E(, k)
v( k v)

p
f
0

(p), (12.23)
The resulting explicit expressions are
K
L
(, k) = 1 +

q
2

2
_
d
3
p

k v
_
k v
k
_
2
1
v

p
f
0

(p), (12.24)
K
T
(, k) = 1 +

q
2

2
_
d
3
p

k v
_
_
v
2

_
k v
k
_
2
_
_
1
v

p
f
0

(p). (12.25)
12.6 Thermal plasma
The next steps in the derivations involve making the non-relativistic approximation p = mv, and
substituting the Maxwellian distribution
f
0

(p) =
n

e
v
2
/2V
2

(2)
3/2
m
3

V
3

. (12.26)
The resulting explicit expressions are
K
L
(, k) = 1

2
p

2
V
2

_
d
3
v

k v
_
k v
k
_
2
e
v
2
/2V
2

(2)
3/2
V
3

, (12.27)
K
T
(, k) = 1

2
p
2
2
V
2

_
d
3
v

k v
_
_
v
2

_
k v
k
_
2
_
_
e
v
2
/2V
2

(2)
3/2
V
3

. (12.28)
The nal step is to carry out the integrals in (12.27) and (12.28). The integrals over d
3
v can be
written in terms of the components of v orthogonal to k and the component along k. The integrals
over the components orthogonal to k are elementary. The integral over the component along k
can be rewritten in terms of a transcendental function called the plasma dispersion function.
110 12. Kinetic theory
12.7 Exercise Set 12
12.1 Let n be an arbitrary unit vector, and let angular brackets denote an average over all
possible orientations of n. To be specic, if n is written in terms of spherical polar coordinates,
n = (sin cos , sin sin , cos ), then the average of any function of n, f(n) say, is
f(n) =
1
4
_
2
0
d
_
1
1
d cos f(n). (12.29)
One has n = 0.
(a) Show that one has
n
i
n
j
=
1
3

ij
. (12.30)
(b) Derive the result 12.30) follows from the following argument. The tensor n
i
n
j
cannot depend
on any vector (because it is an average over all directions) and the only tensorial quantity with
two indices that involves no direction is
ij
. Hence n
i
n
j
must be proportional to
ij
. Writing
n
i
n
j
= A
ij
, show that taking the trace gives A =
1
3
.
(c) Generalize this argument to show
n
i
n
j
n
k
n
l
=
1
15
(
ij
kl +
ik
jl +
il
jk). (12.31)
Chapter 13
Isotropic thermal plasma
Formal treatments of Langmuir waves and of Debye screening are based on the longitudinal re-
sponse function, K
L
(, k), derived using kinetic theory for a thermal distribution of electrons.
13.1 Plasma dispersion function
The Longitudinal and transverse response functions (12.27) and (12.28) can be evaluated in terms
of the plasma dispersion function. One form of this function is referred to as the Fried and Conte
function:
Z(y) =
1/2
_

dt
e
t
2
t y
, (13.1)
A subtle point concerns the evaluation of the integral at the point t = z where the denominator
vanishes. This point is a pole in the integrand, and corresponds physically to the resonance at
= kv. The problem of one deals with this pole was solved by Landau in a paper in 1936, and
the prescription for evaluating around the pole is referred to as the Landau prescription. This
prescription is to give the frequency an innitesimal positive imaginary part. The physical basis
for the Landau prescription is the causal condition: the response must occur at a later time than
the disturbance that causes it. The Landau prescription inside an integral corresponds to writing
1

0
+i0
=
1

0
i(
0
), (13.2)
with
0
= kv here. Equation (13.2) is referred to as the Plemelj formula. The rst term on the
right hand side of (13.2) corresponds to the Cauchy principal value of the integral, and this term
112 13. Isotropic thermal plasma
is real. The other term in (13.2) is imaginary and corresponds to half the residue at the pole,
called the semi-residue. The Plemelj formula is discussed further in B.5.
With y /kv, the Landau prescription corresponds to y y+i0. Applying this prescription
to (13.1), and using (13.2), the real part corresponds to the principal value and the imaginary
part to the semi-residue. The imaginary part is straightforward to evaluate:
ImZ(y) =
1/2
_

dt e
t
2
i(t y) = i

e
y
2
, (13.3)
The real part satises a dierential equation, obtained by dierentiating (13.1) with respect to y
and partially integrating, giving
dZ(y)
dy
= 2[1 +yZ(y)], (13.4)
Writing (y) = yZ(y), one can solve this dierential equation for (y) giving an alternative form
for the real part:
Re (y) = 2y e
y
2
_
y
0
dt e
t
2
. (13.5)
Thus the integral in (13.1) gives
Z(y) = Re
(y)
y
+i
1/2
e
y
2
. (13.6)
Alternatively, one can write (13.6) as
Z(y) = (y)/y, Im(y) = i
1/2
ye
y
2
. (13.7)
The resulting expressions for the longitudinal and transverse parts of the dielectric tensor,
K
L,T
(, k), for a nonrelativistic thermal plasma are
K
L
(, k) = 1 +

2
p
k
2
V
2

[1 (y

) + i
1/2
y

exp(y
2

)], (13.8)
K
T
(, k) = 1 +

2
p
k
2
V
2

[(y

) i
1/2
y

exp(y
2

)], (13.9)
respectively, with
y

= /2
1/2
kV

. (13.10)
Approximate forms are obtained by making expansions of Re (y) for small or large arguments:
Re (y) =
_
_
_
y
2

4
3
y
4
+ for [y
2
[ 1,
1 + (1/2y
2
) + (3/4y
4
) + for [y
2
[ 1.
(13.11)
13.2 Properties of Langmuir waves 113
13.2 Properties of Langmuir waves
The dispersion equation for longitudinal waves follows from (4.23) with (4.22):
K
L
(, k) = 0. (13.12)
When treating Langmuir waves only the contribution of thermal electrons is retained in (13.8).
There are both real and imaginary parts of the response function,
K
L
(, k) = Re K
L
(, k) + iImK
L
(, k), (13.13)
and a solution of (13.12) corresponds to a complex . The real part of the frequency denes the
dispersion relation and the imaginary part describes damping. The absorption coecient,
L
(k),
is equal to 2Im. The solution we seek is of the form =
L
(k) i
L
(k)/2, where L labels
the mode. Provided that the damping is weak, one can solve by initially ignoring the damping,
and then include damping using a perturbation approach. Ignoring the damping, the dispersion
relation is a solution of
Re K
L
(
L
(k), k) = 0. (13.14)
Provided that the damping is weak, the absorption coecient,
L
(k),follows by including the
absorption coecient in the real part, and expanding to nd the imaginary part as a correction:
Re K
L
(
L
(k) i
1
2

L
(k), k) Re K
L
(
L
(k), k) i
1
2

L
(k)
K
L
(, k)

=
L
(k)
, (13.15)
with Re K
L
(
L
(k), k) = 0 corresponding to the solution of (13.14). The absorption coecient
then follows by equating the imaginary parts from (13.12) with (13.13) and (13.15). This gives

L
(k) = 2
L
(k) R
L
(k) ImK
L
(
L
(k), k), (13.16)
where the quantity
R
L
(k) =
_

Re K
L
(, k)

1
=
L
(k)
(13.17)
has a physical interpretation as the ratio of electric to total energy in the waves. The imaginary
part of K
L
(, k) is small in the limit y
2
e
1, which corresponds to phase speeds /[k[ 2
1/2
V
e
,
and this approximate procedure is valid only if this condition is satised.
In the limit y
2
e
1, the contribution of the electrons in (13.8) gives
Re K
L
(, k) 1

2
p

2
_
1 +
3k
2
V
2
e

2
_
, (13.18)
114 13. Isotropic thermal plasma
where the expansion (13.11) is truncated after the term 3/4y
2
. The resulting dispersion relation
for Landmuir waves is

L
(k) (
2
p
+ 3k
2
V
2
e
)
1/2

p
+ 3k
2
V
2
e
/2
p
. (13.19)
Retaining the imaginary parts in (13.16) gives the absorption coecient for Landau damping
of Langmuir waves:

L
(, k) =
_

2
_
1/2

4
L
(k)
k
3
V
3
e
exp
_

2
L
(k)
2k
2
V
2
e
_
. (13.20)
13.3 Ion sound waves
The dispersion equation for ion sound waves is given by (13.12) with contributions from the
electrons approximated assuming y
e
1 and from the ions with y
i
1. These assumptions
require the phase speed of the waves to be in the range V
i
/k V
e
. The approximate form
for K
L
(, k) is
K
L
(, k) 1 +
1
k
2

2
De

2
pi

2
, (13.21)
with
De
= V
e
/
p
, Setting K
L
(, k) = 0 gives
=
s
(k)
kv
s
[1 +k
2

2
De
]
1/2
, v
s
=
pi

De
, (13.22)
where v
s
is the ion sound speed. In the limit k
De
1 the dispersion relation (13.22) reduces to
kv
s
, which is characteristic of a sound wave.
The absorption coecient for ion sound waves is

s
(k)
_

2
_
1/2

s
(k)
_
_
_
v
s
V
e
+
_

s
(k)
kV
i
_
3
e
s(k)
2
/2k
2
V
2
i
_
_
_
. (13.23)
The two terms inside the curly brackets in (13.23) are due to Landau damping by thermal electrons
and by thermal ions, respectively. The damping by thermal ions is strong for v
s
V
i
, and ion
acoustic waves exist as weakly damped waves only for v
s
V
i
, which requires Z
i
T
e
T
i
.
13.4 Debye screening
It follows from (13.8) and (13.11) that the response function K
L
(, k) simplies to 1
2
p
/
2
for
/k and to 1 + 1/k
2

2
D
for /k 0, with
1/
2
D
=

1/
2
D
. (13.24)
13.4 Debye screening 115
The rst of these limits reproduces the cold plasma response. The second corresponds to Debye
screening. This may be seen by comparing a bare Coulomb eld and a screened Coulomb eld in
Fourier space.
Only the static limit, 0, is relevant here, and in this limit only the Fourier transform
in space is relevant. Let (x) be the electrostatic potential associated with a charge q at rest at
the origin r = 0. For a bare charge this corresponds to (x) = q/4
0
r, with r = [x[, and for a
screened charge to (x) = [q/4
0
r] exp(r/
D
). We wish to compare the Fourier counterparts
of these. The spatial Fourier transform is

(k) =
_
d
3
x(x) e
ikx
. (13.25)
One can evaluate the integral as follows. First write the integral in terms of spherical polar
coordinates, r, , ,
d
3
x = d d cos dr r
2
, k x = kr cos . (13.26)
The integral over is trivial and gives 2. The integral over cos give
_
1
1
d cos e
ikr cos
=
2 sin kr
kr
. (13.27)
The nal integral can be written in terms of the standard integral
_

0
dx e
ax
sin mx =
m
m
2
+a
2
, (13.28)
with x r, m k and a = 0 for the bare eld and a 1/
D
for the screened eld. Hence, one
nds

(k) =
_
_
_
q/
0
k
2
for the bare eld,
q/
0
(k
2
+ 1/
2
D
) for the screened eld.
(13.29)
Note that the screened eld is equal to K
L
(0, k) = (1 + 1/k
2

2
D
) times the bare eld.
Before discussing the interpretation of this result further, it is informative to derive it in a
dierent way by starting from Poissons equation and Fourier transforming it. Poissons equation
and its Fourier transform are

2
(x) =
(x)

0
, k
2

(k) =
(k)

0
. (13.30)
The charge density for a charge q at x = x
0
and its Fourier transform are

0
(x) = q
3
(x x
0
),
0
(k) = q e
ikx
0
. (13.31)
116 13. Isotropic thermal plasma
Combining (13.29) and (13.30) gives the eld due to the bare charge:

0
(k) =
q

0
k
2
e
ikx
0
. (13.32)
For x
0
= 0, (13.32) reproduces the result (13.29) derived by Fourier transforming the Coulomb
eld. This conrms that the dependence 1/k
2
is the counterpart in Fourier space of a 1/r
dependence in coordinate space. It also shows that the location of the charge appears only in a
phase factor.
The screening introduces an additional charge density associated with the response of the
plasma to the Coulomb eld due to the charge q. Let this charge density be
ind
(x). The total
charge density associated with the charge q is the sum of the base charge and the induced charge:

tot
(x) = (x) +
ind
(x). The Fourier transform of Poissons equation with
tot
(x) as the source
term gives the screened eld. The result (13.29) for the screened eld shows that one has

tot
(k) = K
L
(0, k)
0
(k), K
L
(0, k) = 1 +
1
k
2

2
D
. (13.33)
More generally, for non-static elds, the self-consistent charge density,
tot
(, k) is given by
K
L
(, k) times the bare charge density,
0
(, k).
The use of the name dielectric constant for K
L
is by analogy with conventional electromag-
netism. The electric eld, E, satises Gauss law with the bare charge density,
0
(x)/
0
, as
the source, and another eld D, called the electric induction, satises Gauss law with
tot
(x) the
source. The phenomenological relation D = E denes the dielectric constant by writing =
0
K.
This corresponds to
tot
= K
0
. In a plasma one needs to distinguish between the longitudinal
and transverse responses, and it is the longitudinal response that is the counterpart of this result
for charges and the elds that they generate.
13.5 Exercise Set 13 117
13.5 Exercise Set 13
13.1 Show that for a nonrelativistic Maxwellian distribution, the integrals that appear in the
general expression (12.19) for K
ij
(], k) may be evaluated in terms of the function
(z) =
z

dt
e
t
2
t z
, (13.34)
Specically, if one writes
_
A(v)
( k v)
a
_
=
1
(2)
3/2
V
3
_
d
3
v
A(v)
( k v)
a
e
v
2
/2V
2
, (13.35)
with z = /

2kV , derive the following results for a = 1:


_
1
k v
_
=
1

2kV
(z)
z
,
_
v
i
k v
_
=
k
i
k
2
[(z) 1], (13.36)
and for a = 2:
_
1
( k v)
2
_
=
1
k
2
V
2
[(z) 1],
_
v
i
( k v)
2
_
=
k
i
k
2
_
1
k
2
V
2
[(z) 1] (z)
_
,
_
v
i
v
j
( k v)
2
_
=

ij
k
2
[(z) 1] +
k
i
k
j
k
4
_
2z
2
[(z) 1] 3(z) + 2
_
. (13.37)
13.2 The approximate dispersion relation for Langmuir wave can be derived from K
L
(, k) = 0
with
K
L
(, k) = 1
2
p
_
1
( k v)
2
_
, (13.38)
where the angular brackets denote an average over a distribution of electrons.
(a) Expand the right hand side of (13.38) in powers of k v/, and show that for an isotropic
distribution of electrons, it gives
K
L
(, k) 1

2
p

2
_
1 +
[k[
2
v
2

2
+
_
, (13.39)
(b) Evaluate v
2
for a Maxwellian distribution and show that K
L
(, k) = 0 implies
2

2
p
+
3[k[
2
V
2
e
.
13.3 Derive the approximation K
L
(, k) 1 + 1/[k[
2

2
D
from (13.38) as follows.
(a) Write the integral over d
3
v in spherical polar coordinates, choosing the polar axis along k.
The integral over azimuthal angle gives 2, and write the integral over cos in the form
_
1
1
d cos
(x cos )
2
=
2
x
2
1
, (13.40)
118 13. Isotropic thermal plasma
with x = /[k[v.
(b) Expand in powers x
2
, retaining only the leading term.
(c) Carry out the integral over v.
13.4 In a solid state plasma the electrons are approximated by a completely degenerate distribution
with occupation number n(p) = 1 below the Fermi energy and n(p) = 0 above the Fermi energy.
The number density of electrons is
n
e
= 2
_
d
3
p
(2 h)
3
n(p), (13.41)
where the factor of two is from summing over the two spin states.
(a) Show that (13.41) implies
n
e
= m
3
e
v
3
F
/3
2
h
3
, (13.42)
where
1
2
m
e
v
2
F
is the Fermi energy and m
e
v
F
is teh Fermi momentum.
(b) Show that in this case one has v
2
= 3v
2
F
/5.
(c) Show that the dispersion relation for Langmuir waves (called plasmons in this case) becomes

2

2
p
+ 3[k[
2
v
2
F
/5. (13.43)
Chapter 14
Plasma instabilities
Plasmas are subject to a variety of dierent instabilities. On a large scale, on which the plasma is
regarded as a magnetized uid, it can be subject to MHD instabilities that disrupt its connement.
Here we are concerned with a dierent kind of instability in which plasma waves grow. These
are called micro-instabilities. The simplest example is when a beam of electrons propagates
through a background plasma and excites Langmuir waves. Even in this case there are two
qualitatively dierent kinds of growth mechanism. One kind of instability is analogous to a maser
or laser, in which absorption is negative, causing waves to grow rather than to damp. The other
kind of instability is physically less obvious. Mathematically, it involves an intrinsically growing
wave. Physically, this may be attributed to a feedback in which the eld of the wave bunches
the particles in phase with the wave, with the eld due to the bunched particles enhancing the
wave eld. These two types of instability are referred to here as maser mechanisms and reactive
mechanisms, respectively. Maser and reactive mechanisms are discussed in this lecture for beam-
driven Langmuir wave, and then the same ideas are applied to some other instabilities.
14.1 Reactive beam instability
Consider a cold electron beam, with velocity v
b
and number density n
b
, moving through a cold
background electron gas with number density n
e
. The dispersion equation for longitudinal waves
is K
L
(, k) = 0, and for this system one has
K
L
(, k) = 1

2
p

2
pb
( k v
b
)
2
, (14.1)
120 14. Plasma instabilities
with
2
p
= e
2
n
e
/
0
m
e
and
2
pb
= e
2
n
b
/
0
m
e
. The denominator in the nal term in (14.1) corre-
sponds to the frequency squared in a frame comoving with the beam. The dispersion equation is
a quartic equation for :
(
2

2
p
)( k v
b
)
2

2
pb

2
= 0. (14.2)
One can always write down the exact solutions of a quartic equation, but it is rarely useful to do so.
(There are two exact solutions for a quadratic equation, three exact solutions for a cubic equation,
and four exact solutions for a quartic equation, but it is not possible to nd exact solutions of
higher order equations in general.) As with the solutions of a quadratic equation, the solutions can
be real or complex. There can be four real solutions, two real solutions and a complex conjugate
pair of solutions, or four complex solutions. We are interested in the case where there are two
real solutions and a complex conjugate pair of solutions. The imaginary part of the frequency for
one of the complex conjugate pair corresponds to a growing wave. The growth of this wave is the
reactive instability.
In the long-wavelength (small k) limit the term k v
b
becomes arbitrarily small, and the four
solutions of (14.2) approach = (
2
p
+
2
pb
)
1/2
, and a double solution at = 0. The solutions
= (
2
p
+
2
pb
)
1/2
correspond to positive and negative frequency solutions for Langmuir waves
in a cold plasma with total number density n
e
+ n
b
. For small k v
b
,= 0 the double solution at
= 0 splits into two real solutions k v
b
; these are called beam modes. As k v
b
increases,
the frequency of one of the beam modes approaches one of the other two solutions, and when
these two coincide, they change into a complex conjugate pair of solutions. Thus the condition
for instability is that there be a double solution of (14.2). The condition for a double solution is

_
(
2

2
p
)( k v
b
)
2

2
pb

2
_
= 0, (14.3)
which is a cubic equation for . The solution of this cubic equation is too cumbersome for semi-
quantitative purposes, and approximations need to be made.
Weak-beam approximation
Several approximations need to be made to derive simple, useful approximations for the growth
rate of the instability. The weak-beam approximation corresponds to n
b
n
e
. Another ap-
proximation is to assume that the growing mode has a frequency close to a beam made. This
14.2 Bump-in-tail instability 121
corresponds to writing
= k v
b
+, (14.4)
and assuming k v
b
. The dispersion equation (14.2) reduces to a suciently simple form in
two cases.
Nonresonant beam mode: For k v
b

p
, (14.2) may be approximated by
()
2
=
n
b
n
e
(k v
b
)
2
. (14.5)
Resonant beam mode: For k v
b

p
, (14.2) with (14.2) may be approximated by
()
3
=
n
b
2n
e

3
p
. (14.6)
Equation (14.5) has two imaginary solutions, one of which corresponds to a growing wave. Equa-
tion (14.5) has three solutions, determined by the three cube roots of unity: 1
1/3
= 1, (1i

3)/2.
One of these solutions corresponds to a growing mode. The growing solutions have
[Im[ =
_
_
_
(n
b
/n
e
)
1/2
[k v
b
[ nonresonant beam mode,
(

3/2

2)(n
b
/n
e
)
1/3

p
resonant beam mode.
, (14.7)
For n
b
n
e
the resonant beam mode has the larger growth rate. Thus, one expects the reactive
instability to occur cause waves with
p
k v
b
to grow, with the growth rate determined
by (14.7) for the resonant beam mode.
14.2 Bump-in-tail instability
The kinetic version of the instability is often called the bump-in-tail instability. The name comes
from assuming that the electron distribution consists of a background Maxwellian distribution
plus an extra distribution of fast electrons, with a mean velocity v
b
. Let the bump-in-tail be
described by a distribution f
b
(p), with p = mv sharply peaked around v = v
b
.
The growth rate for this instability is equal to minus the absorption coecient when the
absorption is negative. The absorption coecient is given in terms of the imaginary part of K
L
by (13.16). The longitudinal part of the dielectric tensor K
ij
, given by (12.19), follows from
K
L
= k
i
k
j
K
ij
/k
2
with. The imaginary part due to the beam follows by imposing the causal
condition on the resonant denominator, and using the Plemelj formula (13.2). This gives
ImK
L
b
(, k) =
e
2

0
k
2
_
d
3
p( k v) k
f
b
(p)
p
. (14.8)
122 14. Plasma instabilities
The absorption coecient due to the beam is given by

L
(k)
p
ImK
L
b
(
p
, k). (14.9)
The simplest treatment of the bump-in-tail instability involves one-dimensional distributions
of streaming electrons and Langmuir waves. Let F(v) be the distribution of electrons integrated
over the velocity components perpendicular to the direction of streaming. Let this one-dimensional
distribution be
n
b
_
dv F
b
(v) =
_
d
3
pf
b
(p). (14.10)
Furthermore, let W(v

)dv

be the energy density in Langmuir waves propagating along the stream-


ing direction with phase speed in the range v

to v

+dv

. The growth of the Langmuir waves is


then described by
dW(v

)
dt
=
L
(v

)W(v

),
L
(v

) =
p
n
b
n
e
v
2

dF(v

)
dv

, (14.11)
where
L
(v

) is the absorption coecient regarded as a function of v

. Maser emission occurs at


v

= v for dF(v)/dv > 0. An idealized example of a bump-in-tail distribution is illustrated in


Fig. 11.2.
Crossover from reactive to kinetic growth
Which of these two instabilities applies? To answer this question one needs to understand the
dierences between the two types of growth. In a reactive instability, the growing wave is coherent
in the sense that it has a well-dened phase. This is because the growth rate is greater than the
bandwidth of the growing waves. Suppose the bandwidth is . Then phase mixing occurs on
a time scale 1/. If the growth rate is greater than , the wave grows faster than phase
mixing occurs, and so it remembers its initial phase. For a kinetic instability, the growth rate
is less than the bandwidth of the growing waves. Phase mixing causes loss of information on
the initial phase before the wave has grown signicantly. This corresponds to the random phase
approximation. The phase is not necessarily random in a maser or laser, it is simply irrelevant to
the growth. Returning to the question: the reactive instability applies if the growth rate exceeds
the bandwidth of the growing waves, and the maser instability applies in the opposite limit.
The bandwidth of the growing waves is determined by the velocity spread of the beam. Let
the velocity spread be v. Then one has F(v) 1/v, dF(v)/dv 1/(v)
2
. Ignoring factors
14.3 Current-driven instabilities 123
of order unity, the growth rate for the resonant reactive instability is
p
(n
b
/n
e
)
1/3
, the growth
rate of the maser instability is
p
(n
b
/n
e
)(v
b
/v)
2
, and the bandwidth of the growing waves is

p
(v/v
b
). One nds that when any two of these are equal, they are also equal to the third.
Hence, one concludes that the reactive instability passes over into the maser instability where the
velocity spread is large enough to make the bandwidth of the growing waves comparable with the
growth rate.
14.3 Current-driven instabilities
A related class of instabilities involves waves driven unstable by a current in the plasma. Once
such an instability occurs, the resulting waves scatter the current carriers, providing a form of
anomalous resistivity. Plasmas in which collisions play no important role are said to be collisionless,
but eective dissipation can still occur in collisionless plasmas due to the anomalous transport
coecient that result from the development of instabilities.
The current density in a plasma is due to the electrons drifting relative to the ions. Consider
the frame in which the ions are at rest, and the electrons are drifting with velocity v
d
. The current
density is then J = en
e
v
d
. In this frame, the electron distribution is
f
e
(p) =
n
2
(2)
3/2
m
3
e
V
3
e
exp
_

(v v
d
)
2
2V
2
e
_
. (14.12)
The absorption coecient for ion sound waves due to thermal particles is given by (13.23). More
generally, when the particles are not necessarily thermal, the absorption coecient is given by

s
(k) =

3
s
(k)

2
pi
ImK
L
(
s
(k), k), (14.13)
where the dispersion relation, =
s
(k), for ion sound waves is given by (13.22). On inserting the
distribution (14.12) into (14.8), the important change to the expression (13.22) for the absorption
coecient is that a multiplicative factor of
s
(k) is replaced by
s
(k) k v
d
. Specically, the
contribution of the electrons to the absorption coecient becomes

s
(k)
_

2
_
1/2
v
s
V
e
(kv
s
k v
d
), (14.14)
where the approximation
s
(k) = kv
s
is made. For k v
d
> kv
s
, the absorption coecient is
negative and ions sound waves grow.
124 14. Plasma instabilities
In practice the waves that grow due to current-driven instabilities are probably not well approx-
imated by ion sound. However, this kind of instability can occur for several dierent low-frequency
wave modes, and should be regarded as a generic mechanism, rather than a specic mechanism for
ion sound waves. It applies to any quasi-longitudinal low-frequency wave mode. The development
of the instability leads to a form of anomalous resistivity irrespective of the details of the grwoing
mode.
14.4 Exercise Set 14 125
14.4 Exercise Set 14
14.1 A plasma consists of thermal electrons and two cold beams of oppositely directed ions. The
beams have equal number densities,
1
2
n
i
, and opposite velocities, v
b
.
(a) Show that for

2kV
e
, the longitudinal part of the dielectric tensor reduces to
K
L
(, k) = 1 +
1
k
2

2
De

2
pi
[
2
+k v
b
)
2
]
[
2
(k v
b
)
2
]
2
. (14.15)
(b) Show that ion sound waves are subject to a reactive instability and determine the maximum
growth rate.
14.2 A plasma with an anisotropic distribution of particles can be subject to several dierent kinds
of instability. The following example is called the Weibel instability, and it causes transverse waves
to grow.
Consider an anisotropic distribution with zero spread in velocity along an axis and a thermal
spread perpendicular to this axis.
f(v) = n(v
z
)
e
v
2

/2V
2

2V
2

. (14.16)
(a) Show that the transverse response function for this distribution is
K
T
(, k) = 1

2
p

2
_
1 +
[k[
2
V
2

2
_
1

2
[k[
2
c
2
__
. (14.17)
(b) Assuming
2
[k[
2
c
2
, show that the dispersion equation for transverse waves becomes

2
(
2
p
+[k[
2
c
2
)
2
p
[k[
2
V
2

= 0. (14.18)
(c) Show that there exist growing solutions of (14.18), and nd the conditions under which the
growth rate is a maximum.
126 14. Plasma instabilities
Appendix A
Cartesian tensors
Tensor algebra, or some equivalent mathematical tool, is needed to describe the response of an
anisotropic medium and to describe stresses and other intrinsically tensorial quantities that appear
in electromagnetic theory. Once mastered, tensor algebra is a useful alternative to vector algebra,
and in many specic examples is simpler to use than vector algebra. The version of tensor algebra
introduced here is for cartesian tensors in three dimensions.
A.1 Vector components and rotations
Consider a vector V. The vector is described by three numbers, which are the components of
the vector along a set of orthonormal basis vectors. Let the basis vectors be e
1
, e
2
, e
3
. The
components of V are
V
i
= V e
i
, with i = 1, 2, 3. (A.1)
The vector itself is written in the form
V =
3

i=1
V
i
e
i
. (A.2)
In vector algebra one denotes a specic vector by a boldface symbol V, and if one wishes to
exhibit its components in a particular coordinate system one writes V = (V
1
, V
2
, V
3
). In tensor
algebra one describes a vector in terms of its components. Thus in the language of tensor algebra
one refers to the vector V
i
. It is implicit that a particular set of basis vectors is chosen, and
that the index i runs over the three components. Here the components are labeled 1, 2, 3, but
they could equally as well be labeled x, y, z, and then i would run over x, y, z rather than 1, 2, 3.
128 A. Cartesian tensors
The index here is denoted by i, but one could choose any other letter, such as j or s, and rewrite
all indices in terms of the new letter without changing the meaning. In principle the labeling of
indices is arbitrary, but it is convenient to restrict it to lower case italic letters as subscripts, and
to use other letters or superscripts to denote labels that are not vector indices. Thus V
Ai
and B
T
j
are interpreted as the ith component of the vector V
A
and the jth component of the vector B
T
respectively.
The vectorial character of a quantity is dened by the way its components transform under
a rotation. It is helpful to distinguish two viewpoints for rotations: these are the active and
passive viewpoints. In an active rotation the coordinate system is held xed and the vector itself
is rotated, and in a passive rotation the vector is held xed and the coordinate axes are rotated.
In an active rotation the vector itself changes. Thus one might describe the initial vector as V and
the vector after rotation as V

. Then if V
i
denotes the vector before the rotation, it is appropriate
to describe the vector after the active rotation by V

i
.
Here only passive rotations are considered. After the passive rotation is performed, let the new
set of basis vectors be e
1
, e
2
, e
3
; note that the primes are added to the indices rather than to the
basis vectors themselves. Then the components of the vector after the rotation are denoted by V
i

with i

running over 1

, 2

, 3

. The vector with respect to the new set of basis vectors is given by
V =
3

=1

V
i
e
i
. (A.3)
A.2 Rotation matrix
A formal way of dening a (passive) rotation is in terms of a rotation matrix. Let the rotation
change the initial set of basis vectors e
1
, e
2
, e
3
into e
1
, e
2
, e
3
. The rotation matrix has components
R
ii
= e
i
e
i
. (A.4)
The relation between the two sets of basis vectors is written
e
i
=
3

=1

(e
i
e
i
) e
i
=
3

=1
R
ii
e
i
, e
i
=
3

i=1
(e
i
e
i
) e
i
=
3

i=1
R
i

i
e
i
. (A.5)
These relations are written in matrix notation:
_
_
_
_
_
e
1
e
2
e
3
_
_
_
_
_
=
_
_
_
_
_
R
11
R
12
R
13

R
21
R
22
R
23

R
31
R
32
R
33

_
_
_
_
_
_
_
_
_
_
e
1

e
2

e
3

_
_
_
_
_
= [R]
_
_
_
_
_
e
1

e
2

e
3

_
_
_
_
_
, (A.6)
A.2 Rotation matrix 129
_
_
_
_
_
e
1

e
2

e
3

_
_
_
_
_
=
_
_
_
_
_
R
1

1
R
1

2
R
1

3
R
2

1
R
2

2
R
2

3
R
3

1
R
3

2
R
3

3
_
_
_
_
_
_
_
_
_
_
e
1
e
2
e
3
_
_
_
_
_
= [R]
T
_
_
_
_
_
e
1
e
2
e
3
_
_
_
_
_
, (A.7)
where [R] denotes the transformation matrix with components R
ii
and [R]
T
denotes its transpose
with components R
i

i
. The matrix [R] is orthogonal, that is, its inverse is equal to its transpose
because the basis vectors e
i
and e
i
are chosen to be orthonormal.
In the formalism of cartesian tensors, a formal denition of a vector is as a set of three
numbers which transform under a rotation in the same way as the basis vectors transform. Thus
if V
i
describes the components of a vector, then under a rotation the components transform to
V
i
=
3

i=1
R
i

i
V
i
. (A.8)
A quantity that is unchanged (invariant) under a rotation is a scalar. A scalar has no index.
A second rank tensor is a quantity that transforms like the outer product of two vectors. Thus a
second rank tensor T
ij
has two indices and under a rotation it transforms to
T
i

j
=
3

i=1
3

j=1
R
i

i
R
j

j
T
ij
. (A.9)
In general a tensor of rank n (for example, T
i
1
i
2
...in
) has n indices (i
1
i
2
. . . i
n
), and it transforms
like the outer product of n vectors. A scalar is a tensor of rank 0, and a vector is a tensor of rank
1.
There are two special tensors whose numerical values are unchanged by a rotation. One of
these is the unit tensor or Kronecker delta:

ij
=
_
1 for i = j,
0 for i ,= j.
(A.10)
The Kronecker delta is dened in terms of the basis vectors by writing

ij
= e
i
e
j
. (A.11)
The orthogonality of the rotation matrix is expressed by
3

=1

R
ii
R
i

j
=
ij
,
3

i=1
R
i

i
R
ij
=
i

j
. (A.12)
The Kronecker delta is a symmetric tensor, that is, it satises
ij
=
ji
.
130 A. Cartesian tensors
The other tensor that has the same value independent of rotations is the permutation symbol

ijk
=
_

_
1 for ijk an even permutation of 123,
1 for ijk an odd permutation of 123,
0 otherwise.
(A.13)
Alternatively the permutation symbol is dened in terms of the vector triple product of the basis
vectors (which are dened to form a right hand set)

ijk
= e
i
e
j
e
k
. (A.14)
The permutation symbol is antisymmetric in all its indices; that is, it changes sign when any two
neighboring indices are interchanged. Thus one has

ijk
=
jik
=
ikj
=
kij
. (A.15)
A.3 Tensor equations
In the foregoing equations, notice that the indices appear either singly or in identical pairs, and
that whenever a pair of identical indices appears the sum over these indices is performed. The
single indices are called free indices, and the pairs of identical indices are called dummy indices.
The number of free indices is equal to the rank of the tensor.
A tensor equation also has a rank. Each element in the tensor equation must have the same
number and kind of free indices. The number of free indices denes the rank of the tensor equation.
An important simplifying assumption adopted in tensor algebra is that the sum over dummy
indices is implied. This summation convention, attributed to Einstein, is that one is to omit the

in all tensor equations, and to note that a sum is included implicitly whenever a pair of identical
indices is present. In a tensor equation the dummy indices are ignored when counting free indices
to determine the rank.
Use of the summation convention allows one to form a scalar from a second rank tensor.
From the tensor T
ij
one may construct the scalar T
ii
, which is just the trace of the tensor (T
ii
=
T
11
+T
22
+T
33
). The proof that this quantity is a scalar follows from the orthogonality relations
(A.12):
T
i

i
= R
i

i
R
i

j
T
ij
= R
ii
R
i

j
T
ij
=
ij
T
ij
= T
ii
, (A.16)
A.4 Rules for tensor equations 131
where the denition (A.4) implies R
i

i
= R
ii
. Note also the use of the Kronecker delta in relabeling
an index:
V
i
=
ij
V
j
. (A.17)
The construction of the scalar T
ii
from the second rank tensor T
ij
is an example of a contraction.
More generally, a contraction refers to the process of setting two free indices equal to each other,
and thus converting them into a pair of dummy indices. This reduces the number of free indices
by two, and so reduces the rank of the tensor or tensor equation by two. Another example of a
contraction is the formation of a vectorial quantity from the outer product of a second rank tensor
and a vector. Such relations appear whenever one describes a response in an anisotropic medium.
For example,
J
i
=
ij
E
j
(A.18)
relates the induced current J to the electric eld E, and where
ij
is the conductivity tensor.
A.4 Rules for tensor equations
The foregoing introduction to cartesian tensors may be summarized and extended in the following
set of rules for writing down tensor equations.
1. Each element in a tensor equation is either a kernel symbol or a product of kernel symbols.
2. The indices are italic subscripts which denote cartesian components, usually understood to
label components 1, 2, 3 or x, y, z. Indices may have aces, such as primes or numerical subscripts,
and two indices are the same only if they have the same ax.
3. Indices appear only either once or twice in any element in a tensor equation. Indices that
appear once are free indices, and indices that appear twice are dummy indices.
4. The summation convention is that the sum (over all cartesian components) over dummy indices
is implied.
5. A tensor equation relates elements each of which has the same number and kind of free indices.
6. The rank of a tensor or of a tensor equation is the number of its free indices. A tensor of rank
zero is a scalar and a tensor of rank one is a vector.
7.The allowed manipulations of a tensor equation include: (i) relabeling free indices, (ii) relabeling
132 A. Cartesian tensors
dummy indices, and (iii) performing contractions (that is, converting two free indices into a pair
of dummy indices).
There are several other denitions that are useful. A second rank tensor T
ij
may be separated
into its symmetric part T
(ij)
and its antisymmetric part T
[ij]
:
T
(ij)
=
1
2
(T
ij
+T
ji
), T
[ij]
=
1
2
(T
ij
T
ji
). (A.19)
A tensor that satises T
ij
= T
ji
is said to be symmetric and a tensor that satises T
ij
=
T
ji
is said to be antisymmetric. For a tensor of rank higher than two one may symmetrize or
antisymmetrize over any two indices. The tensor is said to be completely symmetric only if the
symmetry applies to all pairs of indices, and a tensor is said to be completely antisymmetric only
if the antisymmetry applies to all pairs of neighboring indices.
For a tensor which is complex, it is usually more appropriate to separate into the hermitian
part T
H
ij
and the antihermitian T
A
ij
part:
T
H
ij
=
1
2
(T
ij
+T

ji
), T
A
ij
=
1
2
(T
ij
T

ji
), (A.20)
where the asterisk denotes complex conjugation.
A.5 Vector calculus
The vector calculus is rewritten in tensor notation by noting the basic relations for the scalar and
vector products of two vectors A and B:
A B = A
i
B
i
, (AB)
i
=
ijk
A
j
B
k
. (A.21)
To evaluate a vector triple product one needs to rewrite a product of two permutation symbols
in terms of Kronecker deltas. The basic identity is

abc

ijk
=
ai

bj

ck
+
ak

bi

cj
+
aj

bk

ci

bi

aj

ck

bk

ai

cj

bj

ak

ci
. (A.22)
An important identity follows from (A.22) by making a contraction:

ibc

ijk
=
bj

ck

bk

cj
, (A.23)
where the identities

ab

bc
=
ac
,
ii
= 3 (A.24)
A.5 Vector calculus 133
are used. Performing further contractions on (A.23) leads to the identities

ijc

ijk
= 2
ck
,
ijk

ijk
= 6. (A.25)
The evaluation of the vector triple product involves starting with the second identity (A.21),
which is used twice, and then using (A.23). In this way one nds
[A(BC)]
i
=
ijk
A
j

kab
B
a
C
b
= (
ia

jb

ib

ja
)A
j
B
a
C
b
= A
j
C
j
B
i
A
j
B
j
C
i
= [(A C) B(A B) C]
i
, (A.26)
where the rst identity (A.21) is used in the last step.
In a similar way the identities of the dierential vector calculus are rederived using the tensor
formalism. The basic identities are for grad, div and curl. These involve the dierential operator
= /x, where x is the position vector. One has
(grad)
i
=

x
i
, div A =
A
i
x
i
, (curl A)
i
=
ijk
A
k
x
j
. (A.27)
There are two other operators that appear in the dierential vector calculus:
A grad = A
i

x
i
,
2
=

2
x
i
x
i
. (A.28)
The identities of the dierential vector calculus are:
grad(
1

2
) =
2
grad
1
+
1
grad
2
, (A.29)
div (A) = div A+A grad, (A.30)
curl (A) = curl A+ grad A, (A.31)
curl grad = 0, (A.32)
div curl A = 0, (A.33)
curl curl A = grad div A
2
A, (A.34)
grad (A B) = (A grad)B+ (B grad)A+Acurl B+Bcurl A, (A.35)
div (AB) = B curl AA curl B, (A.36)
curl (AB) = A div B+ (B grad) AB div A(A grad ) B. (A.37)
134 A. Cartesian tensors
Each of these identities may be derived using tensor notation. The following three examples
illustrate the derivations. The proofs of (A.29)(A.31) involve only the identications (A.27) and
the dierentiation of a product, for example,
[curl (A)]
i
=
ijk

x
j
(A
k
) =
ijk

x
j
A
k
+
ijk
A
k
x
j
= (grad A)
i
+(curl A)
i
. (A.38)
The proofs of (A.32) and (A.33) involve a more subtle argument. One has
div curl A =

x
i

ijk
A
k
x
j
=
ijk

2
A
k
x
i
x
j
= 0,
where in the nal step the fact that
ijk
is antisymmetric in jk and the double derivative is
symmetric in jk is used. More generally, the contraction of a symmetric tensor S
jk
= S
kj
with
an antisymmetric tensor A
jk
= A
kj
gives zero. One has S
jk
A
jk
= S
kj
A
jk
= S
jk
A
kj
= S
jk
A
jk
,
where in the rst and last steps the symmetry and antisymmetry properties are used, respectively,
and in the middle step the dummy indices j and k are interchanged. In the proof of (A.34) the
identity (A.23) is used:
(curl curl A)
i
=
ijk

krs

x
j
_
A
s
x
r
_
= (
ir

js

is

jr
)

2
A
s
x
j
x
r
=

x
i
_
A
s
x
s
_


2
A
i
x
s
x
s
=
_
graddiv A
2
A
_
i
. (A.39)
The remaining proofs are somewhat more lengthy and include the dierential operator (A.28).
For example, for (A.37) one has
[curl (AB)]
i
=
ijk

krs
(A
r
B
s
)
x
j
= (
ir

js

is

jr
)
_
A
r
B
s
x
j
+B
s
A
r
x
j
_
= [A div B+ (B grad) AB div A(A grad) B]
i
. (A.40)
Appendix B
Fourier transforms
For many purposes it is both mathematically convenient and physically relevant to describe a phys-
ical quantity in terms of its Fourier transform, rather than in terms of its space-time dependence.
Fourier transformations are introduced here and their general properties are described.
B.1 Fourier integral theorem
The mathematical basis for Fourier transformations is the it Fourier integral theorem:
For any function f(t) such that
1. f(t) is sectionally continuous over < t < ,
2. f(t) is dened as
lim
0
1
2
[f(t
0
+ ) +f(t
0
)] (B.1)
at any point of discontinuity,
3. f(t) is amplitude integrable, that is,
_

dt [f(t)[ < , (B.2)


the following identity holds:
f(t) =
1
2
_

dy
_

dz f(z)e
iy(zt)
. (B.3)
This theorem is used to dene Fourier transforms. For a function f(t) of time t one denes
the Fourier transform by

f() =
_

dt e
it
f(t). (B.4)
136 B. Fourier transforms
Then (B.3) implies that the inverse transform is given by
f(t) =
_

d
2
e
it

f(). (B.5)
The choice of sign in the exponent in (B.4) is a matter of convention, but once this is chosen (B.3)
implies that the opposite sign must appear in the exponent in (B.5). Similarly, it is a matter of
convention as to where the factor 2 in (B.3) is to be included, and the convention adopted in
(B.4) and (B.5) is that the 2 be associated with the inverse transform. The parameter is the
angular frequency (units: s
1
). An alternative convention is to write = 2, where is the
cyclic frequency (units: Hz); the factor 2 then appears in both exponents.
For spatial Fourier transforms it is conventional to choose the opposite sign in the exponent.
Let A(x) be a function of position vector x. Its Fourier transform is dened by

A(k) =
_
d
3
xe
ikx
A(x), (B.6)
where k is the wave vector and where the integral is over all space. The inverse transform is
A(x) =
_
d
3
k
(2)
3
e
ikx

A(k). (B.7)
In cartesian coordinates d
3
x denotes dx dy dz and similarly d
3
k denotes dk
x
dk
y
dk
z
, and all inte-
grals are over the range to .
For a function F(t, x) of both time and position the Fourier transform is dened by combining
(B.4) and (B.6):

F(, k) =
_
dtd
3
xe
i(tkx)
F(t, x). (B.8)
The inverse transform is
F(t, x) =
_
dd
3
k
(2)
4
e
i(tkx)

F(, k). (B.9)
It is convenient to introduce Fourier transforms to solve certain dierential equations. The
reason for this is that dierential operators are replaced by algebraic operators, so that dierential
equations are replaced by algebraic equations. Let an arrow () denote taking a Fourier transform
in time and space: F(t, x)

F(, k). Then we have F(t, x)/t i

F(, k), gradF(t, x)
ik

F(, k), divV(t, x) ik



V(, k), curl V(t, x) ik

V(, k), as indicated by (3.6).
Where no confusion should result, the tilde on the Fourier transformed quantities is omitted
below. The tilde is included to denote that

f is a dierent mathematical function from f. In
B.2 Properties of Fourier Transforms 137
mathematics one uses dierent kernel functions (here

f and f) to denote dierent mathematical
functions, whereas in physics one uses dierent kernel symbols to denote dierent physical quan-
tities. Thus below f is used to refer to a physical quantity either in time and space or in terms
of frequency and wave vector, with the two being distinguished by the arguments of the function.
The tilde is included only if confusion might arise.
B.2 Properties of Fourier Transforms
Three general properties of Fourier transforms are the reality condition, the power theorem and
the convolution theorem.
Reality Condition: If F(t, x) is real then its Fourier transform satises
F

(, k) = F(, k). (B.10)


Power Theorem: If F(t, x) and G(t, x) are real functions, then one has
_
dtd
3
xF(t, x)G(t, x) =
_
dd
3
k
(2)
4
F

(, k)G(, k). (B.11)


Convolution Theorem: The Fourier transform of a product of functions is equal to the convolution
of their Fourier transforms, and the Fourier transform of the convolution of functions is equal to
the product of their Fourier transforms.
The convolution of two functions F(t, x) and G(t, x) is
H(t, x) = F(t, x) G(t, x) =
_
dt

d
3
x

F(t t

, x x

)G(t

, x

). (B.12)
The theorem implies
H(, k) = F(, k)G(, k). (B.13)
The theorem also implies that
J(t, x) = F(t, x)G(t, x) (B.14)
has Fourier transform
J(, k) = F(, k) G(, k) =
_
d

d
3
k

(2)
4
F(

, k k

)G(

, k

). (B.15)
Note that the convolution is dened without any factor 1/2 in terms of the time and space
variables, and with one factor of 1/2 for each variable in Fourier space.
138 B. Fourier transforms
The proof of the reality condition follows by comparing (B.8) and its complex conjugate, with
F

(t, x) = F(t, x) by hypothesis:

(, k) =
_
dtd
3
xe
i(tkx)
F(t, x) =

F(, k). (B.16)
The proofs of the power theorem and the convolution theorem involve the use of Dirac -functions.
The Dirac -function is proportional to the Fourier transforms of unity:
2() =
_

dt e
it
, (2)
3

3
(k) =
_
d
3
xe
ikx
. (B.17)
In cartesian coordinates
3
(k) means (k
x
)(k
y
)(k
z
).
The proof of the power theorem (B.11) involves writing the functions on the left in terms of
their Fourier transforms, and carrying out the integrals using (B.17). For simplicity in writing,
let us give the proof for functions of t only. One has
_
dt f(t)g(t) =
_
dt
_
d
2
e
it
f()
_
d

2
e
i

t
g(

)
=
_
d
2
_
d

2
f()g(

) 2( +

) =
_
d
2
f()g(). (B.18)
The nal step involves using the reality condition to write g() = g

(). Alternatively, on
replacing the variable of integration by , the integrand becomes f

()g(), which is the


form quoted in (B.11).
The proof of the convolution theorem is similar. Once again considering functions of time only,
the Fourier transform of the convolution h(t) =
_
dt

f(t t

)g(t

) is
h() =
_
dt e
it
_
dt

_
d

2
e
i

(tt

)
f(

)
_
d

2
e
i

g(

)
=
_
d

2
_
d

2
f(

)g(

) 2(

) 2(

) = f()g(). (B.19)
The proof generalizes straightforwardly to multi-dimensional Fourier transforms.
B.3 The Dirac -function
The -function (x) was dened originally by Dirac as that function (a) which is equal to zero for
x ,= 0, and (b) whose integral is unity. That is,
(x) =
_
0 for x ,= 0,
for x = 0,
_
dx (x) = 1, (B.20)
B.3 The Dirac -function 139
where the integral is over a range from x < 0 to x > 0. The -function is an even function:
(x) = (x). (B.21)
Suppose that the -function has as its argument a function f(x). Then according to Diracs
denition, [f(x)] is zero except at the zeros of f(x). Suppose that f(x) has n simple zeros at
x = x
1
, x
2
, . . . , x
n
. Then one has
_

dx [f(x)] =
n

i=1
1
[f

(x
i
)[
, (B.22)
with f

(x) = f(x)/x. The proof of (B.22) is as follows. By hypothesis each zero is a simple zero,
and so one has f(x) = 0 and f

(x) ,= 0 at each zero, and furthermore there is a nite separation


between each zero. It follows that the integral in (B.22) separates into n segments each of which
contains only one zero. Consider the segment of the integral that contains the zero at x = x
i
. The
-function is zero except at x = x
i
according to (B.20), and hence the range of integration may
be shrunk to an arbitrarily small range about x
i
: x
i
< x < x
i
+ . Within this range one
makes a Taylor expansion about x = x
i
,
f(x) = (x x
i
)f

(x
i
) + . (B.23)
Suppose that f

(x
i
) is positive. Then if one writes y = (x x
i
)f

(x
i
) the integral reduces to
_
x
i
+
x
i

dx [(x x
i
)f

(x
i
)] =
1
f

(x
i
)
_
dy (y) =
1
f

(x
i
)
, (B.24)
which establishes (B.22) for cases with f

(x
i
) > 0. For f

(x
i
) < 0 one writes y = (x x
i
)[f

(x
i
)[
and uses the property (B.21) to establish the result.
In practice one is usually concerned with integrals that have a nontrivial integrand, that is, with
a function g(x), say, also appearing in the integrand in (B.22). Provided that g(x) is continuous
at each zero of f(x), (B.22) is generalized to
_

dx g(x) [f(x)] =
n

i=1
g(x
i
)
[f

(x
i
)[
. (B.25)
This equation, with f() = and g() = e
it
, is used to show that the inverse transform of
2() is equal to unity, so that the two denitions (B.17) and (B.20) of () are consistent.
140 B. Fourier transforms
B.4 Truncations
The Dirac -function as dened by (B.17) is the Fourier transform of unity. However, the condition
(B.2) for the Fourier integral theorem to apply is clearly violated for f(t) = 1. This implies that the
Fourier transform of unity does not exist as an ordinary function. The -function may be dened
as a generalized function, that is, as a sequence of ordinary functions. Operations involving the
-function are given meaning by evaluating the limit of the operation applied to each member of
the sequence. Loosely, the generalized function may be thought of as the limit of a sequence of
ordinary functions. A variety of dierent sequences may be chosen, and an important property of
a generalized function is that it is independent of the specic sequence chosen to dene it.
A sequence of functions of that denes () is identied as the Fourier transform of a sequence
of functions of t whose limit is unity. One example is the sequence obtained by truncating: the
truncated form of a function f(t) is
f
T
(t) =
_
_
_
f(t) for T/2 < t < T/2,
0 for [t[ > T/2.
(B.26)
With this truncation the integral in (B.2) is necessarily nite and so the Fourier transforms exist
for arbitrarily large but nite T. Another sequence that is used to dene a truncated function
involves truncating with an exponential. Consider
f

(t) = f(t)e
|t|
(B.27)
in the limit 0. For a wide class of functions f(t) that are not themselves amplitude integrable
in the sense (B.2), the function f

(t) is amplitude integrable.


The Fourier transform of the truncated form of the unit function may be used to construct a
sequence of functions that denes the -function. In the case of (B.26) one considers increasingly
large values of T. The -function is then dened as the limit T of the sequence:
() = lim
T
sin(T/2)

. (B.28)
With the choice (B.27) for the truncation, the appropriate sequence is
() = lim
0

(
2
+
2
)
. (B.29)
B.5 Relations involving generalized functions 141
There are two other discontinuous functions whose Fourier transforms dene important gen-
eralized functions. One of these is the step function
H(t) =
_
_
_
1 for t > 0,
0 for t < 0.
(B.30)
The form (B.27) of truncation gives a Fourier transform

() =
_

0
dt e
itt
=
i
+i
. (B.31)
The generalized function is dened by
1
+i0
= lim
0
1
+i
. (B.32)
The other function is the sign function
sgn(t) =
t
[t[
=
_
_
_
1 for t > 0,
1 for t < 0.
(B.33)
The form (B.27) of truncation gives a Fourier transform
sgn

() =
_

0
dt e
itt

_
0

dt e
it+t
=
i
+i
+
i
i
=
2i

2
+
2
. (B.34)
The generalized function in this case is called the Cauchy principal value function:

= lim
0

2
+
2
=
_
_
_
1/ for ,= 0,
0 for = 0.
(B.35)
The sequence (B.35) may be used to dene (1/), but there are alternative sequences that
are equivalent to it. In particular, consider the sequence in which a function is equal to 1/ for
[[ > and is equal to zero for [[ < , with allowed to approach zero. This alternative sequence
inside an integral corresponds to the conventional denition of the Cauchy principal value of the
integral. The equivalence of the two sequences is addressed in Exercise 4.8. More generally, a
variety of sequences may be used to dene each of the three generalized functions considered
here. For formal purposes it is often convenient to choose gaussian functions, as described in
Exercise 4.2.
B.5 Relations involving generalized functions
There is a relation between the unit, step and sign functions:
H(t) =
1
2
[1 + sgn(t)]. (B.36)
142 B. Fourier transforms
The Fourier transform of (B.36) leads to a relation between the generalized functions. Specically,
it implies the Plemelj formula
1
+i0
=
1

i(). (B.37)
The Plemelj formula (B.37) is used in evaluating certain singular integrals.
There are further identities involving the step, unit and sign functions. These include
[sgn(t)]
2
= 1, [H(t)]
2
= H(t), sgn(t)H(t) = H(t). (B.38)
The Fourier transforms of these are evaluated using the convolution theorem. The rst of (B.38)
leads to the identity

2
_

_
1

_

_
1

_
= (). (B.39)
This identity is used in the inverse of Hilbert transforms, which is another class of integral trans-
form (along with Fourier, Laplace and Mellin transforms). The Hilbert transform f
H
() of f()
is dened by
f
H
() =
1

_
1

_
f(

). (B.40)
Then (B.39) implies that the inverse Hilbert transform is given by
f() =
_

f(

)(

) =
1

_
1

_
f
H
(

), (B.41)
where (B.39) and (B.40) are used.
One often wishes to consider the Fourier transform of an idealized physical quantity, such as
a plane wave, which does not vanish at innity. The Fourier transform then does not exist as
an ordinary function. Implicitly one assumes that a truncation is performed, and one uses the
Fourier transforms ignoring the truncation. In practice this leads to a problem only when the
square of a -function appears. The square of the -function needs to be interpreted in terms of
the truncated function. Using the form (B.26) of the truncation, one has
[2()]
2
=
_
T/2
T/2
dt e
it
_
T/2
T/2
dt

e
it

=
1
2
_
T
T
d(t t

)
_
T
T
d(t +t

) e
i(t+t

)
= T 2(), (B.42)
where the double integral is evaluated as outlined in Figure B.1. The limit T is implicit in
(B.42).
An alternative treatment of the square of the -function starts from the form (B.28). Let us
write
D() =
sin
2
(T/2)

2
. (B.43)
B.5 Relations involving generalized functions 143
t-t'=T
t-t'=-T
t+t'=T
t+t'=-T
t'
t
Figure B.1: In evaluating [2()]
2
using (B.42) the integral over the inner square T/2 < t <
T/2, T/2 < t

< T/2 is evaluated using that over the outer (dashed) square T < t t

< T,
T < t +t

< T.
By inspection, one has D(0) = T
2
/4
2
and the ratio D()/D(0) tends to zero for all ,= 0 in the
limit T . Also one has
_

d D() =
T
2
_

dxsinc
2
x =
T
2
, (B.44)
with
sinc x =
sin(x)
x
. (B.45)
Now, as in (B.20), the () is dened as the limit of a sequence of functions each of which is zero
for ,= 0 and each of whose integral is equal to unity. It then follows that D() is just T/2
times (), in accord with (B.42).
The square of the spatial -function is treated in an analogous way. The spatial functions are
assumed to be truncated to an arbitrarily large but nite volume V . Then one has
[(2)
3

3
(k)]
2
= V (2)
3

3
(k). (B.46)
There is an alternative derivation of (B.46) based on truncating to a box. The idea is to use a box
normalization and allow the size of the box to tend to innity. Let the box have sides of length
L
x
, L
y
, L
z
, so that its volume is V = L
x
L
y
L
z
. Allowed solutions have nodes at the sides of the
box, and so correspond to k
x
= 2n
x
/L
x
, k
y
= 2n
y
/L
y
, k
z
= 2n
z
/L
z
with n
x
, n
y
, n
z
arbitrary
144 B. Fourier transforms
integers. The equality of two wavenumbers may be expressed in terms of the Kronecker delta:
for example,
nxn

x
implies equality of two wavenumbers in the x direction. In the limit L
x
,

nxn

x
becomes (2/L
x
)(k
x
k

x
) with k

x
= 2n

x
/L
x
. The Kronecker delta has only the values 0
and 1, each of which is equal to its square, and hence the Kronecker delta satises the identity
[
nxn

x
]
2
=
nxn

x
. (B.47)
In the limit L
x
this identity implies
[(2/L
x
)(k
x
k

x
)]
2
= (2/L
x
)(k
x
k

x
). (B.48)
Taking the product of this latter identity with the corresponding identities in the y and z compo-
nents, the result reduces to (B.46) with k replaced by k k

.
B.6 Alternative treatment of Plemelj formula
To show
lim
0+
_
b
a
dx
f(x)
z +i
= T
_
b
a
dx
f(x)
x
i f(0). (B.49)
Separate the integral into three parts
_
b
a
dx
f(x)
z +i
=
_

a
dx
f(x)
x +i
+
_

dx
f(x)
x +i
+
_
b

dx
f(x)
z +i
(B.50)
with . Denote the three parts by J
1
, J
2
, J
3
, respectively. One has
lim
0+
/
[J
1
+J
3
] = lim
0+
_
_

a
bdx
f(x)
x
+
_
b

dx
f(x)
x
_
= T
_
b
a
dx
f(x)
x
, (B.51)
which is the denition of the Cauchy principal value of the integral. The remaining term gives
lim
0+
/
J
2
= lim
0+
/
_

dx
f(0)
x +i
= lim
0+
/
_
_

dx
f(0)x
x
2
+
2
i
_

dx
f(0)
x
2
+
2
_
= lim
0+
/
[2if(0) arctan(/)] = if(0). (B.52)
Appendix C
Assignment Sets
C.1 Assignment Set 1: Due 20 May 2011
E1.1 Calculate the plasma frequency,
p
, the electron cyclotron frequency,
e
= eB/m
e
, the Debye
length,
D
, the Alfven speed, v
A
, and the collision frequency,
c
, for the following parameters:
(a) A laboratory fusion plasma: n
e
= 10
15
m
3
, B = 1 T, T
e
= 10 eV.
(b) The ionosphere: n
e
= 0.1 m
3
, B = 10
5
T, T
e
= 10
3
K.
(c) The solar corona: n
e
= 10
10
cm
3
, B = 1 G, T
e
= 10
6
K.
(d) The interplanetary medium: n
e
= 1 cm
3
, B = 3 10
6
G, T
e
= 10
6
K.
E1.2 The dispersion relation for Langmuir waves may be approximated by
2
L
(k) =
2
p
+ 3k
2
V
2
e
,
where V
e
is the thermal speed of electrons. Show the the phase speed, v

, and the group speed,


v
g
, for waves with this dispersion relation satisfy v

v
g
= 3V
2
e
.
E1.3 A plasma consists of electrons and protons with equal number densities, n, with a uniform
magnetic eld along the x-axis with z the vertical direction.
(a) Calculate the current density, J, due to the gravitational drift of electrons and protons, where
the gravitational acceleration, g, is along the negative z-axis.
(b) Calculate the force per unit volume, J B, due to this current density.
(c) Give a physical interpretation of your answer.
E1.4 The group velocity may be evaluated in spherical polar coordinates using
v
gM
=
c
(n
M
)/
_

1
n
M
n
M

t
_
(C.1)
146 C. Assignment Sets
with = (sin , 0, cos ), t = (cos , 0, sin ), with the magnetic eld along b = (0, 01).
Evaluate the group velocity for Alfven waves assuming = kv
A
[ cos [ and hence n
A
=
c/v
A
[ cos [. Hence show that the group velocity is v
gA
= v
A
(0, 0, cos /[ cos [).
E1.5 The magnetoionic waves are nearly circularly polarized and nearly linearly polarized in two
opposite limits depending on which term dominates in the expression (7.5) for
2
.
(a) Show that the angle at which the two contributions are equal is

0
= arcsin
_
2
1/2
Y
[(1 X)
4
+Y
2
(1 X)
2
]
1/2
(1 X)
2

1/2
_
. (C.2)
(b) Show that (E13.9) reduces to
0
arccos
1
2
Y for X 1, Y 1.
(c) Estimate the range of angles for which the approximation of circular polarization breaks down
for 1 GHz radiation in the fully ionized region of the ISM.
C.2 Assignment Set 2: Due 3 June 2011 147
C.2 Assignment Set 2: Due 3 June 2011
E2.1A (Advanced stream only) Collisions are included in the longitudinal response function
in the cold plasma limit by including a drag term in the equation of motion for the electrons:
m
e
dv
dt
= eE
c
m
e
v. (C.3)
Repeat the cold plasma derivation of the dielectric constant in (4.20), showing that it gives
K
L
() = 1

2
p
( +i
c
)
. (C.4)
Identify the imaginary part and use (11.4), viz.

L
(k) = 2
L
(k)R
L
(k) ImK
L
(
L
(k), k), (C.5)
to calculate the absorption coecient due to collisional damping for transverse waves.
E2.2A (Advanced stream only) A plasma consists of thermal electrons and two cold beams
of oppositely directed ions. The beams have equal number densities,
1
2
n
i
, and opposite velocities,
v
b
.
(a) Show that for

2kV
e
, the longitudinal part of the dielectric tensor reduces to
K
L
(, k) = 1 +
1
k
2

2
De

2
pi
[
2
+k v
b
)
2
]
[
2
k v
b
)
2
]
2
. (C.6)
(b) Show that ion sound waves are subject to a reactive instability and determine the maximum
growth rate.
E2.1N (Normal stream only) In magnetron discharge (RF discharge with magnetic eld), the
plasma has the electron number density 3 10
15
m
3
and electron temperature 5 eV. The total
density of anions (negatively charged ions) is quarter of the electron number density. The gas in
discharge vessel which can be treated as an ideal gas has pressure 5 Pa, and temperature 300 K.
The magnitude of magnetic eld is 0.01 T (assume that the magnetic eld is uniform). For this
plasma, calculate:
a) The total density of positive ions. What assumption do you make?
148 C. Assignment Sets
b) The electron Debye screening length, and the total screening length (given the anion temper-
ature 0.1 eV, and positive ion temperature 0.05 eV) .
c) The total number density of neutrals in the gas discharge and the ionization degree of this
plasma.
d) The plasma frequency, ion-neutral collision frequency and mean free path for a 2 eV anion.
Assume that the collision cross-section of the anion with neutrals is = 5 10
15
cm
2
and
the mass is 18 amu.
e) The cyclotron frequency and Larmor radius of an electron with an energy of 5 eV; what is the
direction of gyration of the electron?
f) The cyclotron frequency and Larmor radius of a positive ion (mass of a ion is 24 amu) with an
energy of 0.5 eV, what is direction of gyration of the ion?
E2.2N (Normal stream only) In fusion plasma under certain conditions Coulomb collisions
appear to be dominant.
(a) What can you say about the ionization degree of this plasma? For given electron n
e
and
neutral n
n
number densities, and ion-neutral collision cross-section
in
nd the electron
temperature under which the ion-electrons collision frequency is equal to the ion-neutral
collision frequency (assume the charge of ion +1e).
(b) List methods of plasma connement. Explain the balance of what forces and pressures makes
possible to conne the plasma cylinder by an axial current? Name at least two problems
with the plasma connement in the case considered.
(c) In experiment with magnetized plasmas, you are aiming at increasing the ion temperature
without signicantly aecting the electron temperature. Which would you choose among
the following waves?
(i) Alfven wave
(ii) Left-Hand Circularly-Polarized (LHCP)
(iii) Helicon wave
C.2 Assignment Set 2: Due 3 June 2011 149
(iv) Ordinary (O) wave
(v) Extraordinary (X) wave
(vi) Right-Hand Circularly-Polarized (RHCP)
Please explain your reasons.
150 C. Assignment Sets

Das könnte Ihnen auch gefallen