Main
Main
Preface vi
I Introduction 1
1 Electromagnetic dispersion 2
1.1 Basic equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.1 Maxwell’s equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Electrodynamics in a linear medium . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.3 Dispersion operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Waves in homogeneous linear media . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.1 Basic concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.2 Dispersion relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.3 Quasimonochromatic waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
i
CONTENTS ii
ii
CONTENTS iii
iii
CONTENTS iv
iv
CONTENTS v
Appendices 154
A Abbreviations 154
B Notation 155
C Conventions 158
C.1 Numbers and matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
C.2 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
Bibliography 160
v
Preface
Plasmas support a wide variety of waves. These waves significantly determine plasma dynamics, and
they are also indispensable for plasma manipulation and diagnostics.
Studying plasma waves is an essential part of studying plasma physics and its applications. It is also
useful for understanding waves in general, because the complexity of waves in plasmas demands a
particularly systematic approach to wave theory.
This course is intended as an introduction into physics of (mostly linear) plasma waves and briefly
covers the following general topics:
• concept of linear dispersion;
• dispersion operators and their symbols;
• geometrical-optics approximation;
• envelope equation, ray tracing, mode conversion;
• transport of the wave action, energy, and momentum;
• dispersion properties of nonmagnetized and magnetized plasma within fluid and kinetic models;
• basic types of plasma waves and their applications to plasma manipulation and diagnostics;
• basic instabilities and mechanisms of collisionless dissipation;
• nonlinear saturation of kinetic instabilities, elements of quasilinear theory.
For in-depth discussions of these topics, see additional literature, for example, Refs. [1–8].
vi
Part I
Introduction
The purpose of this first, intentionally haphazard, part of the course is to familiarize
readers with some basic vocabulary of plasma-wave theory. The concepts introduced in
this part will be used later for developing a more systematic theory.
1
Lecture 1
Electromagnetic dispersion
where E and B are the electric and magnetic field, respectively, j is the current density, ρ is the
charge density, and c is the speed of light. (Gaussian units will be used throughout the course.)
Magnetic Gauss’s law can be considered as an initial condition for Faraday’s law, because
∂t (∇ · B) = ∇ · ∂t B = −c∇ · (∇ × E) = 0, (1.2)
where we used the fact that the divergence of a curl is identically zero. Similarly, electric Gauss’s law
can be considered as an initial condition for Ampere’s law, because
∂t ρ + ∇ · j = 0. (1.4)
In this sense, not all of Eqs. (1.1) are entirely independent from each other, and it will be sufficient
for us to use only a subset of them for studying waves. We will return to this subject later.
2
LECTURE 1. ELECTROMAGNETIC DISPERSION 3
The “induced” current density j (i) is determined by medium’s response to the electric field, and the
linear operator σ̂ that determines this response is called conductivity.2 The remaining, “free” current
density, j (f) , is independent of E; it includes the current density that is prescribed externally, and
it can also include a (generally time-dependent) current density that is determined by the initial
conditions. For this reason, specifying the representation (1.5) requires specifying the initial moment
of time since which the dynamics is considered. We will denote this moment as t0 ; then, by definition,
Below, we consider j (f) as prescribed and discuss how to model the induced current density j (i) .
In some cases, the latter can be as simple as
where σ is some matrix function or even a scalar. This model is commonly known as Ohm’s law. It
is also called the local-response model, because it assumes that the current at a given location (t, x)
is determined by the field only at the same location (t, x). Such an approximation can be reasonable,
for example, for modeling strongly collisional media. However, in general, j (i) (t, x) can also depend
on E(t0 , x0 ) at (t0 , x0 ) other than (t, x). Such media are called dispersive. Nonlocality in time is called
temporal dispersion, and nonlocality in space is called spatial dispersion.
The general induced current in a dispersive medium can be expressed as a functional
t ∞
j (i) (t, x) = d t0 d x0 Σ(t, x, t0 , x0 )E(t0 , x0 ). (1.8)
t0 −∞
The tensor Σ(t, x, t0 , x0 ), which is the “coordinate” (time-space) representation of σ̂ (see also Box 1.1),
can be formally defined as a functional derivative
δj(t, x)
Σ(t, x, t0 , x0 ) = . (1.9)
δE(t0 , x0 )
In other words, it serves as the weight function that determines how much the field E(t0 , x0 ) contributes
to the current j (i) (t, x). The fact that the integration domain in Eq. (1.10) is limited to t0 < t reflects
causality; that is, the current at a given time t can be affected by past fields (t0 < t) but not by future
fields (t0 > t). One can also replace the upper integration limit t in Eq. (1.8) with ∞,
∞ ∞
j (i) (t, x) = d t0 d x0 Σ(t, x, t0 , x0 )E(t0 , x0 ), (1.10)
t0 −∞
ω̂ = i ∂t , k̂ = −i ∇,
. .
(1.11)
1 We do not involve B here because for oscillatory fields that we are interested in, B can always be expressed through E
using Faraday’s law. For stationary fields, the general form of the induced current density is j (i) = σ̂E + κ̂B.
2 We use caret ˆ to denote integral operators, including differential operators as a special case.
3
LECTURE 1. ELECTROMAGNETIC DISPERSION 4
The current density j (i) is also often expressed in the following alternative form:
∞ ∞
t + t0 x + x0
(i)
j (t, x) = dt 0
d x Σ̄ 2 , 2 , t − t , x − x E(t0 , x0 ),
0 0 0
t0 −∞
where the “symmetrized” kernel Σ̄ is connected with the previously introduced Σ as follows:
.
where = denotes definitions. In particular, using Eq. (1.4), one can express ρ through E via ω̂:
Using this notation, one can write the solution of Eq. (1.12) as
where ρ(f) may include contributions from initial conditions and j (f) but is independent of E. It is
also convenient to introduce the susceptibility operator
4π i
χ̂ = σ̂ (1.15)
ω̂
(we assume the notation 1/ω̂ ≡ ω̂ −1 ), which we also represent as
t ∞
(χ̂E)(t, x) = d t0 d x0 X (t, x, t0 , x0 )E(t0 , x0 ), (1.16)
t0 −∞
ˆ = 1̂ + χ̂. (1.17)
where 1̂ is a unit matrix operator. Then, Eq. (1.14) can be written as ρ = ρ(f) + ρ(i) . Here, the
“induced” charge density is given by
ρ(i) = −∇ · P (1.18)
and P is the so-called electric polarization, which is understood as the electric dipole moment per
unit volume [9] and is defined as3
. χ̂
P = E. (1.19)
4π
3 In this course, we define χ̂ using Stix’s notation [1]. Sometimes, though, the factor 4π is absorbed in the definition
.
of χ̂; then, ˆ
= 1̂ + 4π χ̂ and P = χ̂E.
4
LECTURE 1. ELECTROMAGNETIC DISPERSION 5
Also, the field ˆE = E +4πP is called the electric displacement field. With these, Maxwell’s equations
(1.1) can be re-written as follows:
1 ∂ 2 (ˆE) 4π ∂j (f)
∇×∇×E+ 2 2
=− 2 . (1.21)
c ∂t c ∂t
We will limit our consideration to regions where the external currents are zero. This does not auto-
matically mean that j (f) is zero, because the initial conditions can give rise to time-dependent currents
(Sec. 1.1.2), including currents resonant to waves of interest. Such currents can significantly affect
the wave propagation. However, we will postpone a detailed discussion of this subject until Part IV.
For the time being, let us simply assume that: (i) the integral (1.8) that defines j (i) converges at
t0 → −∞ and (ii) the initial conditions at t0 → −∞ do not affect waves of interest at finite t. One
can expect this to be a reasonable assumption, for example, in collisional media, because collisions
destroy information about the initial conditions naturally. Also, if the wave amplitude grows, then
j (i) is determined by E mainly from the recent past and, again, the initial conditions at t0 → −∞
should not matter. Hence, we adopt t0 → −∞, so σ̂E is given by the following integral that converges
by our assumption:
t ∞
(σ̂E)(t, x) = d t0 d x0 Σ(t, x, t0 , x0 )E(t0 , x0 ), (1.22)
−∞ −∞
. c2
D̂ E E = 0, D̂ E = 2 (k̂k̂† − 1 k̂ 2 ) + ˆ. (1.23)
ω̂
The operator D̂ E will be called the (electromagnetic) dispersion operator.
Note that other representations of the wave equation are also possible. For example, it can be
convenient to use the electrostatic potential (Problem PI.1) or the magnetic field (Sec. 2.3.3) instead
of E, or even give up the very notion of the dielectric operator (Problem PI.2).
4 Global modes, which are described by discrete variables ψ(t), are also waves in the sense that ψ(t) can be considered
5
LECTURE 1. ELECTROMAGNETIC DISPERSION 6
X (t, x, t0 , x0 ) = X̄ (t − t0 , x − x0 ). (1.25)
Accordingly, Σ̄ab (t, x) can be understood as the ath component of the current, ja (t, x), produced by
the bth component of the electric field of the form6 Eb (t0 , x0 ) = δ(t0 − 0)δ(x0 ).
It is easy to see from here that homogeneous linear media support monochromatic waves, that is,
waves of the form (Box 1.2)
where the amplitude E , the frequency ω, and the wavevector k are complex constants. Indeed, in this
case, Eq. (1.26) leads to
∞ ∞
(σ̂E)(t, x) = dτ d s Σ̄(τ, s)E e −i ω(t−τ )+i k·(x−s) (1.28)
0 −∞
∞ ∞
= dτ d s Σ̄(τ, s) e i ωτ −i k·s E(t, x). (1.29)
0 −∞
This function is called the spectral representation of σ̂, or simply the spectral conductivity, because
σ(ω, k) is obtained from Σ̄ by applying the Fourier transform in space and the Laplace transform
in time. The physical meaning of this quantity can be understood by noticing that according to
5 To put it differently, the function Σ̄ introduced in Box 1.1 is independent of t̄ and x̄.
6 The added 0 in the argument of the delta functiondenotes an infinitesimal positive value. This shift ensures that
the
t time integral over
the domain (t0 , t) is well defined, tt d t0 δ(t − t0 − 0)= 0t−t0 d τ δ(τ − 0)= −∞∞
d τ δ(τ ) = 1, while
t
0
d t0 δ(t − t0 )= ∞ d τ δ(τ ) is undefined. In our special 0case, t → −∞, but this is not essential.
0 0
6
LECTURE 1. ELECTROMAGNETIC DISPERSION 7
Although the actual electric field is real, the sourceless wave equations that we work with allow
solutions in a complex form. These equations have real coefficients, so for any complex field E c
that is a solution (the index c stands for “complex”), the complex-conjugate field E ∗c is a solution
too, and so is any linear combination of the two. Hence, one can search for real E in the form
E = Re E c = 1/2 (E c + E ∗c ).
To simplify the notation, we will not distinguish E and E c where it is not essential.
Eq. (1.26), j is a convolution of Σ̄ and E; thus, its spectral representation j(ω, k) is simply propor-
tional to the spectral representation E(ω, k) of the electric field with the proportionality coefficient
σ(ω, k):7
j(ω, k) = σ(ω, k)E(ω, k). (1.31)
Since Σ̄(t, x) = 0 at t < 0, the integral over t in Eq. (1.30) can be formally extended to −∞,
∞ ∞
σ(ω, k) = dt d x Σ̄(t, x) e i ωt−i k·x . (1.32)
−∞ −∞
Still, it should not be confused with the Fourier transform, because ω is generally complex. Since
|Σ̄(t, x)e −i ωt+i k·x | = |Σ̄(t, x)|e −ω t , either ωi = Im ω should be sufficiently large or the response func-
i
.
tion Σ̄(t, x) should diminish rapidly enough at t → ∞ for the integral (1.30) to converge. In particular,
if Σ̄(t, x) ∝ e −νt with constant ν, then the integrals (1.30) and (1.32) exist provided that
ωi + ν > 0. (1.33)
These are the same requirements as those adopted in Sec. 1.1.3; that is, applicability of our theory is
facilitated by collisions and by exponential growth of the wave amplitude. See also Appendix AI.1 for
basic properties of response functions like σ.
From the spectral representation of σ̂, one readily finds the spectral representation of ˆ, which is
known as the dielectric tensor :
(ω, k) = 1 + χ(ω, k), (1.34)
with 1 being the unit matrix. Here, χ is the spectral representation of χ̂; it is defined by analogy
with Eq. (1.30) and given by
4π i
χ(ω, k) = σ(ω, k), (1.35)
ω
as shown in Box 1.3. Then, the wave equation (1.23) can be written as follows:8
D E (ω, k)E = 0, D E (ω, k) = N N † − 1 N 2 + (ω, k). (1.36)
The matrix D E is called a dispersion matrix or dispersion tensor, and N is the refractive-index vector:
.
N = ck/ω. (1.37)
7 See theory of the Fourier and Laplace transforms. We will revisit this topic in Part IV, where we will also discuss
objects in inhomogeneous media are operators, because their representations can be defined in more than one way and
can differ significantly. For example, “the dielectric tensor” of inhomogeneous plasma is undefined until a convention is
specified for the mapping ˆ → (t, x, ω, k). Except in special settings, such as cold stationary plasmas, one should not
expect (t, x, ω, k) to be a meaningful quantity if it is obtained simply by taking (ω, k) of homogeneous plasma and
replacing the constant plasma parameters with functions of (t, x). We will revisit this issue in Part II (Box 3.2).
7
LECTURE 1. ELECTROMAGNETIC DISPERSION 8
For t = t0 , this gives F (t0 , x) = 4πj (i) (t0 , x) = 0 for any E; thus, one also has X̄ (0, x) = 0 at
all x. Then, the above equation can be written as
t ∞
4π(σ̂E)(t, x) = dt 0
d x0 [∂t X̄ (t − t0 , x − x0 )]E(t0 , x0 ),
t0 −∞
which means that ∂t X̄ = 4π Σ̄. Then, from Eq. (1.30), one finds
∞ ∞
4πσ(ω, k) = dt d x [∂t X̄ (t, x)] e i ωt−i k·x = i ωχ(ω, k),
0 −∞
where we used integration by parts and X̄ (0, x) = 0. This leads to Eq. (1.35).
or simply D = 0 when D = D is scalar. This can be considered as an equation for ω(k), which is called
a dispersion relation. The solutions for ω at given k, ω = ωq (k), are called dispersion branches. Note
that they are generally complex, and we will assume the following notation throughout the course:
ω = ωr + i ωi ,
. .
ωr = Re ω, ωi = Im ω. (1.40)
Depending on the number of these branches b,9 or the order of the wave equation, the waves are
attributed as scalar waves (b = 1) or as vector waves (b > 1); see also Box 1.4.
A spatially monochromatic field with a given wavevector k can contain contributions from multiple
branches, called (eigen)modes, and can be expressed as follows:10
ψ̄ q e −i ωq (k)t+i k·x ,
X
ψ(t, x) = (1.41)
q
9 Integral
wave equations generally yield infinitely many branches and will be discussed in Part IV.
10 When solutions for ω are degenerate, it is in principle possible to have solutions ∝ ta e −i ωq (k)t+i k·x with natural a.
However, this is not typical for waves of our interest, so we will not consider such cases.
8
LECTURE 1. ELECTROMAGNETIC DISPERSION 9
Any differential equation ∂tm ψ = F[ψ, ∂t ψ, . . . ∂tm−1 ψ] (where F may be an integral transform
in x) can be represented as ∂t ψ̃ = F̄(ψ̃), where ψ̃ = (ψ̃ 1 , . . . ψ̃ m−1 )| is a vector of dimension
. . .
m × dim ψ, with ψ̃ 1 = ψ, ψ̃ 2 = ∂t ψ, . . . , ψ̃ m = ∂tm−1 ψ̃. Assuming F is linear, such an equa-
tion generally has b = m × dim ψ roots, so it describes vector waves unless ψ is scalar and m = 1.
For example, consider the Schrödinger equation and the Klein–Gordon equation from quantum
mechanics. (As will be discussed later, they also emerge in plasma-wave theory.) The Schrödinger
equation has dim ψ = m = 1, so b = 1; i.e., it has only one dispersion branch and thus the
corresponding waves are true scalar waves. The Klein–Gordon equation also has dim ψ = 1; but
in this case, m = 2, so b = 2, and the corresponding waves are vector waves. The representation
of the Klein–Gordon equation as a first-order equation for a two-dimensional vector is known as
the Feshbach–Villars representation. A similar reduction of plasma-wave equations to first-order
vector equations is discussed, for example, in Problem PI.2.
where the coefficients ψ̄ q are determined by the initial conditions. Each of these coefficients must
satisfy D[ωq (k), k]ψ̄ q = 0. Thus, it can be expressed as ψ̄ q = Aq hq (k), where Aq is a scalar amplitude
and hq (k) is a unit polarization vector 11 defined via
In general, dim hq = dim ψ can be different from dim x (Problem PI.2), in which case hq may not
have a clear physical meaning. But let us mention the important special case when ψ is a complexified
(Box 1.2) single-mode electric field on a three-dimensional space, with the real field being
If hq is parallel to some fixed coordinate axis a (hq = ēa , where ēa is the unit vector along the a
axis), such a field√ is called
√linearly polarized, because E remains parallel to the a axis at all times. If
hq = (ēa ±i ēb )/ 2 (here 2 is added only to ensure the normalization |hq | = 1), then Ea2 +Eb2 = const.
This means that E at given x rotates in the (a, b) plane with a constant amplitude, √ so the electric
field is called circularly polarized in this plane. Similarly, if hq = (ēa + i ς ēb )/ 2 with general real ς,
the trajectory of E in the (a, b) plane at fixed x is an ellipse, so the field is called elliptically polarized.
9
LECTURE 1. ELECTROMAGNETIC DISPERSION 10
Suppose a quasimonochromatic wave, i.e., a wave such that ψ(k) is localized around some k = k0 ,
.
so only small κ = k − k0 contribute to the integral (1.45). For simplicity, let us assume that ωi is
small enough, so we can expand ω(k) as follows:
ω = ω0 + i γ + v g · κ, (1.46)
. .
where ω0 = ωr (k0 ), γ = ωi (k0 ), and v g is the group velocity given by
. ∂ωr (k)
vg = . (1.47)
∂k k=k0
This leads to
ψ(t, x) = e i k0 ·(x−vp t)+γt
∞
d κ ψ̄(k + κ) e i κ·(x−vg t) , (1.48)
0
−∞ (2π)n
where we introduced the “phase velocity”
. k0 ω0
vp = . (1.49)
k0 k0
(The corresponding phase speed is vp = ω/k.) This can also be written equivalently as
Exercise 1.1: Show that if the first-order expansion of ωr in κ [Eq. (1.46)] is replaced with
the second-order expansion of ωr in κ, then Eq. (1.52) acquires a form similar to the Schrödinger
equation of a free quantum particle,
The above discussion shows that dispersion relations ω(k) contain information not only about
monochromatic oscillations but also about the evolution of wave packets. Because of this, dispersion
relations are particularly important and will be the primary focus of this course.
10
LECTURE 1. ELECTROMAGNETIC DISPERSION 11
11
Lecture 2
In this lecture, we illustrate the key concepts introduced in Lecture 1 and also give a sneak preview
of the generic transformations of waves in inhomogeneous plasmas, which will be discussed more
systematically in Part II. To do this, we describe waves in cold nonmagnetized plasmas as a simple
example. We will also discuss these waves later in a broader context.
Here, the summation is taken over all species s, es are charges of these species (e.g., for electrons,
ee = −e < 0), ns are the corresponding densities, and v s are the corresponding flow velocities. Tildes
denote that the corresponding quantities are small linear perturbations, while background currents
and fields are denoted with index 0. For simplicity, we assume that there are no background flows,
v 0s = 0, so ns0 may not depend on t but may depend on x. In this case,
X X X
j = j̃ = es n0s ṽ s + es ñṽ s ≈ es ns0 ṽ s , (2.2)
s s s
Let us assume for now that there is no background magnetic field (B 0 = 0); then, the magnetic part
of the Lorentz force is quadratic in Ẽ and thus can be neglected. The term (ṽ s · ∇) ṽ s is negligible for
the same reason. Let us also assume that the pressure term ∇Ps can be neglected too. (The validity
conditions of this approximation will become clear when we study kinetic theory.) For the collision
12
LECTURE 2. A SNEAK PREVIEW: WAVES IN COLD NONMAGNETIZED PLASMA 13
term C s , we adopt a simple model C s = −νs ṽ s , where νs = νs (x) serves as the collision rate. Then,
Eq. (2.3) becomes
∂ ṽ s es
= −νs ṽ s + Ẽ. (2.4)
∂t ms
This leads to the following equation for the current density:
2
∂ j̃ s ωps
= −νs j̃ s + Ẽ, (2.5)
∂t 4π
where we introduced the so-called plasma frequencies
s
. 4πns0 e2s
ωps = . (2.6)
ms
i ωps
2 (f)
∂ j̃ s
j̃ s = j̃ (f)
s + Ẽ, = −νs j̃ (f)
s . (2.7)
4π(ω̂ + i νs ) ∂t
Since the total current density is a sum of j̃ s , one can introduce the conductivity σ̂ s for each species
independently, with the total conductivity σ̂ being the sum over those of the individual species,
X
σ̂ = σ̂ s . (2.8)
s
Also note that in our problem, all these operators are scalar operators, σ̂ s = σ̂ s , and the same applies
to χ̂ and ˆ. In summary then, an inhomogeneous stationary cold nonmagnetized plasma without
average flows is characterized by the following operators:
i ωps
2 2
ωps X 2
ωps
σ̂ s = , χ̂s = − , ˆ = 1 − . (2.9)
4π(ω̂ + i νs ) ω̂(ω̂ + i νs ) s
ω̂(ω̂ + i νs )
In order to find the operators (ω̂ + i νs )−1 explicitly, let us directly integrate Eq. (2.5). This leads
to (Exercise 2.1)
2 t
ωps
s (t, x) = j̃ s (t0 , x) e d t0 e ν (x)(t −t) Ẽ(t0 , x).
0
−νs (x)(t−t0 )
j̃ (f) (f)
, j̃ (i)
s (t, x) =
s
(2.10)
4π t0
From here, we can readily infer the coordinate representation (1.24) of the conductivity operator:
2
ωps (x) −νs (x)(t−t0 )
Σs = Σs , Σs (t, x, t0 , x0 ) =
4π
e δ(x − x0 )H (t − t0 ), (2.11)
13
LECTURE 2. A SNEAK PREVIEW: WAVES IN COLD NONMAGNETIZED PLASMA 14
where we have added the Heaviside step function H (t − t0 ) to emphasize that Σ(t, x, t0 , x0 ) ≡ 0 for
t0 > t. The presence of the delta function in Eq. (2.11) signifies that the plasma response is local in
space, so there is no spatial dispersion.1
The effect of the conductivity operator on the electric field can now be written as
t ∞ 2
ωps (x) −νs (x)(t−t0 )
(σ̂ s E)(t, x) = dt 0
d x0 4π
e δ(x − x0 )E(t0 , x0 )
t0 −∞
2 t
ωps (x)
=
4π
d t0 e −ν (x)(t−t ) E(t0 , x).
s
0
(2.12)
t0
For fields monochromatic in time with frequency ω > −iνs , this yields
i ωps
2
(x)
(σ̂ s E)(t, x) = E(t, x). (2.13)
4π(ω + i νs (x))
Because x enters here only as a parameter, one can introduce the local conductivities, the local
susceptibilities, and the local dielectric tensor as functions of (ω; x):
i ωps
2 2
ωps X 2
ωps
σs (ω; x) = , χs (ω; x) = − , (ω; x) = 1 − , (2.14)
4π(ω + i νs ) ω(ω + i νs ) s
ω(ω + i νs )
2
where the dependence on x enters through ωps and νs .
14
LECTURE 2. A SNEAK PREVIEW: WAVES IN COLD NONMAGNETIZED PLASMA 15
Then, (ω) = 0 leads to a quadratic equation for ω and thus has two solutions, while (ω) = N 2 leads
to a cubic equation for ω and thus has three solutions, so there are five modes overall. Although exact
solutions for ω(k) are possible in both cases, it is more instructive to consider the regime when νe is
small. By assuming ω νe and justifying this assumption a posteriori one obtains
ω ≈ ±ωpe −
i νe , (2.19a)
2
q
2 + c2 k 2 −
ω ≈ ± ωpe
i νe
2
ωpe
. (2.19b)
2 c2 k 2 + ωpe
2
c2 k 2
ω ≈ −i νe . (2.19c)
c2 k 2 + ωpe
2
Because Eq. (2.16) has exactly five roots, no solutions other solutions are possible.
Note that all five modes (2.19) satisfy ωi > −νe . Thus, the applicability condition of our theory,
νe + ωi > 0, is satisfied and the result (2.19) can be trusted. One might wonder if this happened
by accident, and the answer is no. As seen from Eq. (1.8), Σ̄ is just the current density produced
by a field that is delta-shaped in time (and in space). In contrast, the solutions (2.19) describe
collective oscillations, in which particles remain coupled with nonzero self-consistent field at all t. As
collisions dissipate the electron energy, this loss is partially compensated from the energy of the field,
so the current dissipates slower than electrons would slow down without the wave. This argument
also extends to general collisional plasmas, so in such plasmas one can always obtain σ(ω, k) from σ̂
simply by replacing ω̂ with ω (assuming that the plasma parameters are time-independent). However,
this is not the case in collisionless plasmas,2 where the total macroscopic field can dissipate while the
response of individual particles does not. For the time being, we will always assume some amount
of collisional dissipation to ensure the applicability of our general approach; i.e., collisionless plasmas
will be considered as collisional plasmas with infinitesimally small but positive νs :
νs → 0 + . (2.20)
In what follows, Eq. (2.20) will always be assumed for simplicity, in which case is simplified as
ωp2
s
. X
2 .
(ω) = 1 − 2 , ωp = ωps (2.21)
ω s
“The” plasma frequency ωp absorbs information about both electrons and ions, so we do not have
to neglect the ion response to treat dispersion relations analytically. Then the resulting three wave
modes are as follows.
nutshell, (relevant) wave modes are defined differently when the collision rates are small enough, and then, in a way,
small collision rates cease to matter. We will return to this subject in the second half of the course.
15
LECTURE 2. A SNEAK PREVIEW: WAVES IN COLD NONMAGNETIZED PLASMA 16
Exercise 2.2: Derive a PDE for the electron-density perturbations in this wave using the
linearized continuity equation for the electron density, the linearized momentum equation for the
electron velocity, and the electrostatic dispersion relation derived in Lecture 1.
In electron–ion plasma, one has ωp ≈ ωpe (due to mi me ), in which case the plasma oscillations
can be interpreted as oscillations of electrons relative to the stationary neutralizing ion background.
These oscillations were discovered by Langmuir and Tonks in the 1920s [11], so they are commonly
called electron Langmuir oscillations. The terms is also often shortened to just “Langmuir oscillations”,
but note that ion Langmuir oscillations are possible too. They occur at ωp ≈ ωpi when electrons are
hot enough to have negligible susceptibility. We will discuss this subject in Lecture 7.
16
LECTURE 2. A SNEAK PREVIEW: WAVES IN COLD NONMAGNETIZED PLASMA 17
d 2E + q(x)E = 0, (2.27)
d x2
where we have adopted the notation Ẽ = Re [hE (x)e −i ωt ] and introduced
. ω2 ω 2 − ωp2 (x)
q(x) = 2 (ω; x) = . (2.28)
c c2
Now that the coefficients of the wave equation explicitly depend on x, a spatially monochromatic
wave is no longer a solution. But let us assume (until Sec. 2.3.2) that the plasma parameters change
along x slowly, the condition for which is yet to be specified. Then locally, q can be considered
√
constant, so q serves as the local wavenumber k. [This can be seen by comparing Eq. (2.28) with
the homogeneous-plasma dispersion relation (2.24) or simply by solving Eq. (2.27) locally under the
approximation of constant q.] This interpretation is justified provided that k is reasonably well defined,
.
i.e., if the corresponding wavelength λ = 2π/k has a small gradient (which is a dimensionless quantity):
d λ 1, 2π
λ= √ . (2.29)
dx q
Then, it is possible to construct an asymptotic solution of Eq. (2.27) using the so-called Wentzel–
Kramers–Brillouin (WKB) approximation. We will not review the general WKB method here, but
will rather present an ad hoc solution of Eq. (2.27) sufficient for our purposes. (In Part II, we will
adopt an alternative approach, which is easier to apply in the general case.)
√
Let ε 1 be the corresponding characteristic value of d λ/d x ∼ (Lc q)−1 , where Lc is some
constant characteristic scale of the plasma density. Then, Eq. (2.27) can be written as
ε2 E 00 + Q(ξ)E = 0, (2.30)
where the prime denotes d /d ξ, ξ = x/Lc , and Q(ξ) = (εLc )2 q(x) is an order-one function with
. .
Q0 ∼ Q. Let us search for a solution in the form E = e i S(ξ)/ε , where S is a complex function of the
form S = S0 + εS1 + ε2 S2 . . . with order-one Sn . Then, Eq. (2.30) yields3
17
LECTURE 2. A SNEAK PREVIEW: WAVES IN COLD NONMAGNETIZED PLASMA 18
Since the equality is supposed to hold for all small ε, this leads to
and the terms Sn>1 will be ignored because including them changes E only by O(ε). Then,
S0 = ± dξ
p
Q(ξ), S1 =
i ln Q(ξ) + const, (2.33)
4
so the field can be expressed as follows:
exp ± i dx
A p
E (x) = p
4
q(x) , (2.34)
q(x)
where A is an arbitrary complex constant and ± correspond to the waves propagating in the ±x
directions, respectively. Since Eq. (2.27) is linear, any linear combination of the solutions with plus
and minus sign is also a solution.
The WKB solution (2.34) corroborates our expectation that the local wavevector k(x), defined
as the gradient of the phase, ispthe same4 as that predicted by the homogeneous-plasma dispersion
relation (2.24); i.e., k(x) = ± q(x). Equation (2.34) also shows the field amplitude changes as
|E | ∝ [q(x)]−1/4 , so
p
|E (x)|2 q(x) = const. (2.35)
This coincides with the well-known adiabatic invariant of a harmonic oscillator [our original Eq. (2.27)
is the equation of a harmonic oscillator], which is the ratio of the oscillator’s energy and frequency [12].
A more general form of this conservation law will be discussed in Lecture 5.
Exercise 2.3: Show that the time-average Poynting vector of an evanescent wave with a single
imaginary wavevector is zero. Explain how waves can carry electromagnetic energy across a
finite-width region where they are evanescent.
Let us assume a linear approximation for ωp2 (x) and choose coordinates such that the cutoff cor-
responds to x = 0, with waves propagating at x < 0 and evanescent at x > 0. Then locally,
x
ωp2 (x) = ω 2 1 + , Lc > 0. (2.36)
Lc
d 2E
1/3
Lc c2
. x .
− η E = 0, η= , `= . (2.37)
d η2 ` ω2
4 Although this is always true approximately, we will show in Lecture 4 that a small deviation from the homogeneous-
18
LECTURE 2. A SNEAK PREVIEW: WAVES IN COLD NONMAGNETIZED PLASMA 19
E
0.8
0.6
0.4
0.2
η
-20 -15 -10 -5
-0.2
-0.4
Figure 2.1: The Airy function Ai (η) (blue) vs. the WKB approximation (2.40) (red), with x = η`,
q = −η`2 , and C = 1. The field is a standing wave at η < 0 and an evanescent wave at η > 0.
This is the so-called Airy equation, and its general solution is a linear superposition of the Airy
functions of the first and second kind, Ai (η) and Bi (η), respectively. Assuming that a wave is
incident from the left, we are interested in a solution that corresponds to E (x → ∞) = 0. This leaves
us with E = CAi (η), where C is an arbitrary constant that determines the maximum near-cutoff
magnitude, E max ∼ C (Fig. 2.1). We can connect C with the constants A+ and A− of the general
WKB solution,
A e i θ(x) + A− e −i θ(x) x
E (x) = +
π
d x0
p
, θ(x) = + q(x0 ), (2.38)
[q(x)]1/4 4 0
by comparing its asymptotic behavior with the known asymptotic behavior of Ai (η) at η → −∞:
cos[θA (η)] π 2
Ai (η) ≈ √ , θA (η) = − (−η)3/2 . (2.39)
π (−η)1/4 4 3
Near the cutoff, q ≈ −η/`2 , which leads to θ(x) ≈ θA (η). Then, for the
√ solution (2.38) to match
E = CAi (η) at not-too-large negative η, one should adopt A+ = A− = 2 π`C, which leads to
C cos[θ(x)]
E (x) = √ . (2.40)
π[q(x)`2 ]1/4
In other words, the WKB approximation
p predicts that at large negative η, the field is a standing wave
with the local wavenumber k(x) = q(x) [same as predicted by the homogeneous-plasma dispersion
relation (2.24)] and the amplitude |E | that satisfies
1/6
|E |
1 1 c
∼ = √ . (2.41)
E max (q`2 )1/4 4
ωLc
Since the right-hand side depends on the plasma parameters slowly, the field amplification near the
cutoff is not as dramatic as one might imagine based on the WKB approximation. Nevertheless, the
WKB approximation describes the field well all the way up to η ∼ −1 (Fig. 2.1).
19
LECTURE 2. A SNEAK PREVIEW: WAVES IN COLD NONMAGNETIZED PLASMA 20
In principle, the WKB approximation can be reinstated near a cutoff by changing the represen-
tation of the field equation. For example, consider the Airy equation (2.37) in the operator form,
k̂ 2 E + x̂E = 0. In the usual, coordinate representation, one has x̂ = x and k̂ = −i ∂x . (Here, we
blur the distinction between the dimensionless η and −i ∂η on one side and the dimensional x and
−i ∂x on the other side for simplicity.) However, we can adopt the wavevector representation (cf.
momentum representation in quantum mechanics), i.e., take the Fourier transform of the Airy
equation. Then, x̂ = i ∂k and k̂ = k, so the equation becomes
k 2 Ē + i ∂k Ē = 0,
where Ē is the Fourier spectrum of E . This new equation can be solved using the WKB
method. In fact, the corresponding WKB approximation, Ē = const × exp(i k 3 /3), is an exact
solution.
Mapping it back to the coordinate space is only a matter of calculating the integral
d k exp(i kx + i k3 /3), which can be done at various levels of accuracy.
One can also use a more general class of transforms called metaplectic transforms. The latter
can rotate (more generally, perform symplectic transformations of) the “phase space” (x, k)
such that in the new phase space (x0 , k 0 ), the wavenumber k 0 is never zero and thus cutoffs are
eliminated [13] (Fig. 2.2). The Fourier transform can be considered as a particular case of the
metaplectic transform that corresponds to 90◦ rotation.
TE wave
Let us start with the transverse-electric (TE) polarization (Fig. 2.3), which corresponds to E x = E y =
0. Since ∂z E z = 0, one has ∇ · Ẽ = 0, which leads to the following equation for E z :
0 = ∇2 E z +
ω2
(ω; x)E z =
d 2 E z + 1 [ω2 − c2 k2 − ω2 (x)]E . (2.43)
c2 d x2 c2 y p z
This equation is similar to Eq. (2.27) up to replacing ω 2 → ω 2 − c2 ky2 . Thus, the field forms the same
WKB/Airy profile (Fig. 2.1), except the cutoff is shifted to the location where ωp2 (x) = ω 2 − ky2 c2 , in
agreement with the homogeneous-plasma dispersion relation (2.24).
TM wave
Let us also consider the transverse-magnetic (TM) polarization (Fig. 2.3). In this case, ∇ · Ẽ is
nonzero, which means that a transverse wave is coupled to the oscillations of the charge density, ρ̃,
through Gauss’s law (1.1c). Similarly, ρ̃ is coupled to Ẽ via (Exercise 2.4)
e2
∂t2 ρ̃ + ωp2 ρ̃ = f˜, f˜ = − ∇n0e · Ẽ. (2.44)
me
Equation (2.44) can be understood as the electrostatic wave equation for free Langmuir oscillations
driven by an external force f˜ ∝ Ẽ. If this force has a resonant frequency ω = ±ωp , the magnitude of
ρ̃ can become large. Then, several scenarios are possible.
20
LECTURE 2. A SNEAK PREVIEW: WAVES IN COLD NONMAGNETIZED PLASMA 21
Figure 2.2: An illustration of cutoff elimination using the metaplectic transform (Box 2.1): the phase
space is continuously transformed such that the wavenumber remains constant and thus a wave never
sees a cutoff in this transforming frame. In the simplest case, the transformation is a rotation by some
nonzero angle α, with α = 90◦ corresponding to the Fourier transform.
Figure 2.3: A schematic of transverse-wave oblique incidence on plasma with background electron
density n0e = n0e (x) illustrating the difference between a TE wave and a TM wave. The vertical red
line indicates the critical-density region, where ωp2 (x) = ω 2 . The dashed red line indicates the cutoff
region, where ωp2 (x) = ω 2 − ky2 c2 . Waves are propagating on the left and evanescent on the right.
The background color intensity illustrates the background density, with darker colors corresponding
to higher densities.
21
LECTURE 2. A SNEAK PREVIEW: WAVES IN COLD NONMAGNETIZED PLASMA 22
Ex
x
0
Figure 2.4: A typical E x [Eq. (2.46), numerical solution] for intermediate β and weak dissipation,
with the boundary conditions corresponding to vanishing B z at x → ∞. Here, (ω; x) is given by
Eq. (2.18), with ωp2 (x) given by Eq. (2.36).
When the angle β between k and ∇n0e approaches zero, one recovers the usual WKB/Airy profile
discussed above. Likewise, if β is large, one can expect from Eq. (2.24) that the wave is reflected
much earlier than it reaches the critical-density region. Then, f˜ is extremely small in the resonance
region, so Langmuir oscillations are not excited (unless the collision rate is also extremely small) and
a TM wave behaves just like a TE wave. However, at intermediate β, the coupling of a TM wave with
Langmuir oscillations can be substantial and results in a significant peak of the wave field near the
critical region. This can be described by the following equations (Exercise 2.4)
d 2 B z − ∂ ln (ω; x) d B z + ω2 (ω; x) − k2 B = 0,
(2.45)
d x2 ∂x dx c2 y z
ck B
Ex = − y z , Ey =
ic dBz . (2.46)
ω(ω; x) ω(ω; x) d x
A typical E x [Eq. (2.46)] for intermediate β is shown in Fig. 2.4, where the Airy pattern is seen on the
left and a large near-singular field is seen near the Langmuir resonance. (For details and asymptotic
analytic solutions, see Ref. [14].) Note that a small collision rate has to be added to keep the wave
field finite. This is due to the fact that energy is continuously deposited by the incoming wave into
Langmuir oscillations and cannot escape the critical-density region, because Langmuir waves are tied
to this region and because they cannot dissipate collisionlessly in cold plasma. One can show that
adding thermal effects removes both these limitations.
The coupling of TM waves with Langmuir oscillations is an example of mode conversion, which is
mathematically similar to quantum tunneling. Note that this effect is completely missed within the
homogeneous-plasma analysis (Sec. 2.2), where the two transverse modes have identical properties
and are completely uncoupled from Langmuir oscillations. To capture polarization effects robustly
and in a general geometry, a more systematic theory is needed. Later, we will present such theory,
which will be readily applicable to almost any dispersion operators.
22
Appendices for Part I
Let us assume that our system is stable in the sense that a response to a delta-shaped field eventually
fades away, i.e., Σ̄(t → +∞) = 0.5 Then, the integral that determines its Laplace image,
∞
σ(ω) = d t e i ωt Σ̄(t), (2.48)
0
converges for all ω that satisfy Im ω ≥ 0. Similarly, if Σ̄(t) fades away faster than any power of t at
t → ∞ (which is typically the case, as discussed in Lecture 8), then all integrals of the form
∞
.
Jn [Σ̄](ω) = d t e i ωt (i t)n Σ̄(t) (2.49)
0
converge too. [The square brackets denote that Jn is a functional of Σ̄, and (ω) denotes that Jn
depends on ω, as usual.] Since Jn [Σ̄](ω) = σ (n) (ω), where (n) is the nth derivative, this means that
all derivatives of σ(ω) are well defined at Im ω ≥ 0. This proves the following theorem:
response of a current to an external field, which is described by the conductivity operator σ̂, is typically stable. (Systems
exhibiting avalanche ionization are exceptions.) In contrast, the response of a self-consistent field to an external current,
which is described by D̂ −1
E , is often not.
23
APPENDICES FOR PART I 24
[Here, we have assumed that the function Σ̄ is also sufficiently well behaved so that Σ̄(0) and J[Σ̄0 ] are
finite.] The function Σ̄(t) is determined by microscopic processes that have some nonzero minimum
time scale T . Then, roughly, |J[Σ̄0 ]| . |J[Σ̄]|/T . Hence, at ω T −1 , the second term on the right-
hand side of Eq. (2.50) can be neglected compared to the term on the left-hand side. Thus, at large
enough ω, one has σ(ω) ≈ i Σ̄(0)/ω, and in particular,
σ(ω → ∞) = 0. (2.51)
In combination with Eq. (2.51), the analyticity of σ(ω) in the upper half of the complex-ω plane
(proven above) leads to some interesting properties of the functions
. .
σr (ω) = Re σ(ω), σi (ω) = Im σ(ω). (2.52)
These properties are derived as follows. Let us consider real ω0 and the integral
.
I(ω0 ) = d ω ωσ(ω)
−ω
, (2.53)
C 0
where C is a contour that goes along the real axis and encircles the pole at ω = ω0 from above. On
one hand, one can rewrite I(ω0 ) as
∞
I(ω0 ) = d ω ωσ(ω)
−ω
− i πσ(ω0 ), (2.54)
−∞ 0
where denotes the Cauchy principal value of the corresponding integral. On the other hand, one
can close the contour C through a semicircle C+ with radius R → ∞ at Im ω > 0, because the integral
over C+ is zero. [This is because at ω → ∞, the integrand decreases faster than 1/ω due to σ(ω) → 0.]
But σ(ω)/(ω − ω0 ) is analytic everywhere within the closed contour, so the integral over this closed
contour must be zero. Therefore, I(ω0 ) = 0, which means
σ(ω0 ) = −
i ∞
d ω ωσ(ω) . (2.55)
π −∞ −ω 0
By taking the real and imaginary parts of Eq. (2.55), one obtains that σr and σi are connected via
the Hilbert transform; namely,
∞
σr (ω0 ) =
1
π
d ω ωσ−
i (ω)
ω
, (2.56a)
−∞ 0
∞
σi (ω0 ) = −
1
π
d ω ωσr−(ω)
ω
. (2.56b)
−∞ 0
These are known as the Kramers–Kronig relations. They show that having nonzero σi at some frequen-
cies implies having nonzero σr at some (possibly, different) frequencies, and vice versa. Notably, the
collisionless dissipation discussed in Part IV can be anticipated from the Kramers–Kronig relations.
24
Problems for Part I
(a) Consider the components of E parallel and perpendicular to the wavevector, E k and E ⊥ . Using
Eq. (1.36), estimate the ratio of |E k | and |E ⊥ | at sufficiently large N and argue that
(b) Assuming that the field is electrostatic and that the medium is described by an unspecified
dielectric tensor , derive the corresponding dispersion relation from Gauss’s law. Show that the
same result is obtained from Eq. (1.36) if one takes for granted that the field is electrostatic.
(c) In a gyrotropic medium with no spatial dispersion, the dielectric tensor has the form
⊥ (ω) −i g(ω)
0
(ω, k) = i g(ω) ⊥ (ω) 0 . (2.58)
0 0 k (ω)
Consider an electrostatic wave propagating with a such medium with k = {k⊥ , 0, kk }. (One can
always choose axes such that ky = 0.) Substitute this k and Eq. (2.58) for into the dispersion
relation derived in problem (b). Using your result, show that the group velocity of such a wave
is orthogonal to the wave phase velocity.
(a) Show that the combination of the linearized momentum equation for the fluid velocities ṽ s ,
Ampere’s law for the wave electric field Ẽ, and Faraday’s law for the wave magnetic field Ẽ can
be represented together in the form
i ∂t ψ = Ĥψ, (2.59)
25
PROBLEMS FOR PART I 26
i ωp1 (x)
−α · Ω1 (x) 0 ... 0 0
0 −α · Ω2 (x) . . . 0 i ωp2 (x) 0
.. .. .. .. .. ..
Ĥ =
. . . . . . .
(2.60)
0 0 . . . −α · Ω N (x) i ω pN (x) 0
−i ωp1 (x) −i ωp2 (x) . . . −i ωpN (x) i cα · k̂
0
0 0 ... 0 −i cα · k̂ 0
. . p
Here, Ωs = es B 0 (x)/(ms c), ωps = es 4πn0s (x)/ms (note that this definition is slightly differ-
.
ent from the one that we used earlier), and α = (αx , αy , αz )| is the column vector comprised of
the following Hermitian matrices:6
0 0 i 0 −i 0
0 0 0
0 0 −i , αy = 0 0 0 , αz = i
. . .
αx = 0 0 . (2.61)
0 i 0 −i 0 0 0 0 0
Note that Eq. (2.59) is similar to the Schrödinger equation for vector particles. (This becomes
even more evident if one multiplies it by ~ and expresses the right-hand side through the mo-
.
mentum operator p̂ = ~k̂ instead of the wavevector operator k̂.) Thus, ψ can be understood as
the photon wave function in cold plasma (cf. the photon wave function in vacuum [15]).
Hint: Use ψ = (ζ̃ 1 , ζ̃ 2 , . . . , ζ̃ N , Ẽ, B̃)| , where ζ̃ s is a rescaled v s , with the rescaling
factor that you are asked to find. Also use that for any three-component column
vectors A and B, one has A×B = −i (α·A)B, as can be verified by direct calculation.
2 2
(b) Calculate |ψ|
, which is 2the same as ψ here. What is the physical meaning of this quantity?
Show that d x |ψ(t, x)| ≡ hψ|ψi is conserved.
Hint: In the last question, use Eq. (2.59) and the fact that Ĥ is Hermitian. Do not
use the explicit formula (2.60), or your calculations will be much longer than necessary.
(c) Argue that cold-plasma modes governed by Eq. (2.59) cannot be unstable, irrespective of how
inhomogeneous the plasma and B 0 are.7
(a) Consider bulk electrons and beam electrons as different species. Show that the beam-electron
.
susceptibility in the spectral representation is given by χb (ω, k) = −ωb2 /(ω − kvb )2 , where ωb =
p
4πnb e2 /me . (You may adopt ω̂ = ω and k̂ = k.)
6 Notably, αa belong to the family of Gell–Mann matrices, which serve as infinitesimal generators of SU(3).
7 This is true only for cold plasma without flows. Otherwise, Ĥ is different and generally has different properties.
26
PROBLEMS FOR PART I 27
(b) Calculate the plasma dielectric function (ω, k) and write down the general dispersion relation
of electrostatic oscillations.
(c) How many branches does this dispersion relation have?
(d) Show graphically, by qualitatively analyzing the function (ω, k) derived in (b) and using your
answer from (c), that such plasma is unstable at small enough k.
(e) For the unstable regime, plot Re ω and Im ω as functions of k. (You may do it qualitatively or
by solving the dispersion relation numerically, but make sure to plot all branches.)
.
(f) Assuming η = nb /n0 is a small parameter, show that the maximum of the growth rate γ scales
as γmax ∝ η 1/3 . (If a beam is warm, which case will be studied later in this course, the scaling
for γ can be different.)
Hint: Use the fact that, to the zeroth order in η, the unstable branch has ω = kvb
p from (e)]. Show that to the leading order, γ → ∞ at
[as you should be able to see
2 .
k = ±ωp0 /vb , where ωp0 = 4πn0 e2 /me . Hence, adopt k = ±ωp0 /vb as the optimum
wavenumber for the instability and then find γ more accurately.
Hint: Remember that the tangential component of the electric field is always continuous
at the plasma boundary. Make sure you understand why.
(a) Using Maxwell’s equations for the three components of the electric field Ẽ and the three com-
ponents of the magnetic field B̃, show that the above assumptions lead to Ẽθ = B̃r = B̃z = 0
at r 6= a. (Assume that 6= 0.) Argue that in this problem, B̃θ is continuous at the plasma
boundary.
(b) Show that ∇ · Ẽ = 0 at r 6= a. Using this, show that E z satisfies the modified Bessel equation
d 2 E z + 1 d E z − E = 0. (2.62)
d ζ2 ζ d ζ z
. .
Here ζ = κr, κ2 = kz2 +ωp2 /c2 −ω 2 /c2 , ωp (r < a) = ωp0 is a nonzero constant, and ωp (r > a) = 0.
Using the continuity of Ẽz at r = a, find E z (r) as a piecewise-analytic function. (Remember
that your solution is supposed to be finite everywhere, including r → 0 and r → ∞.)
(c) Using the continuity of B̃θ at r = a (which can be expressed through d E z /d ζ), show that the
dispersion relation is
ωp2
1 I1 (κin a) 1 K1 (κout a)
1− 2 + = 0. (2.63)
ω κin a I0 (κin a) κout a K0 (κout a)
Here, In and Kn are modified Bessel functions of the first and second kind, respectively; also,
. .
κin = κ(r < a) and κout = κ(r > a).
.
(d) Solve Eq. (2.63) numerically. Plot ω(kz ) and E z (r) for several values of a/δc , where δp = c/ωp0 .
Explain the results qualitatively.
27
Part II
Basic theory of
quasimonochromatic waves
28
Lecture 3
In this lecture, we derive an approximate envelope equation for a general quasimonochromatic wave
by asymptotically expanding the wave dispersion operator. In doing so, we also introduce the Wigner–
Weyl transform, which is a central element of modern wave theory and will be used also in other parts
of the course.
theories include the quasioptical approximation and the paraxial approximation as a special case (Box 3.1).
29
LECTURE 3. ASYMPTOTIC EXPANSION OF DISPERSION OPERATORS 30
3.2 Notation
To shorten the calculations, we will describe the wave propagation in terms of spacetime coordi-
.
nates xα , or in the invariant notation, x = {x0 , x}, where x0 = ct and xa with a > 0 are spatial
.
coordinates. The dimension of the coordinate space n = dim x can be any n ≥ 0. The spacetime
.
dimension is n = dim x = n + 1 and restricted to n ≥ 1. For simplicity, we restrict our consideration
to spacetimes Mn with the Minkowski metric gαβ = g αβ = diag {−1, 1, 1, . . .}.3 Then, w̄ and k̄ can be
expressed through the components of the row vector
.
k̄α = ∂α θ = (−w̄/c, k̄)α , (3.6)
.
where ∂α = ∂/∂xα ; in particular, k̄0 = −w̄/c. The corresponding column vector has components
k = g kβ , where g αβ is the inverse metric. In case of the Minkowski metric, g αβ = gαβ , so
α αβ
k 0 = −k̄0 = w̄/c and k̄ a = k̄a , or in the invariant form, k̄ = (w̄/c, k̄)| . This implies
The operators ω̂ = i ∂t and k̂ = −i ∇ that we introduced earlier can be similarly expressed through
. .
k̂ α = −i ∂α .
.
(3.8)
By analogy with quantum mechanics, let us consider scalar fields on Mn as vectors |ψi in the
corresponding Hilbert space. Then, any scalar function ψ(x) can be understood as the x-representation
of the corresponding vector |ψi; namely, ψ(x0 ) = hx0 |ψi. Here, |x0 i is the normalized eigenvector of
the coordinate operator x̂ that corresponds to the eigenvalue x0 ; i.e.,
.
where M(x, x0 ) = hx|M̂ |x0 i is the x-representation of M̂ . Similarly, one can introduce the normalized
eigenvectors of the wavevector operator k̂,
and define the k-representation of vectors as hk|ψi and of operators as hk|M̂ |k0 i (Exercise 3.1).
2 Reported here is a modern approach to GO [16, 17]. Earlier formulations of GO [18] do not use the term WWT
30
LECTURE 3. ASYMPTOTIC EXPANSION OF DISPERSION OPERATORS 31
The inverse WWT4 W −1 maps any given function M (x, k) to an operator M̂ via [19]
M̂ =
1
(2π)n
d x d k d s |x + s/2i M (x, k) e i k·s hx − s/2| . (3.17)
For M̂ that is a matrix of operators M̂ uv , the corresponding symbol M (x, k) is defined as the matrix
whose elements are the symbols of M̂ uv . Also note that the WWT commutes with the Hermitian
conjugation:
W [M̂ † ] = M † (x, k). (3.18)
4 The direct WWT is also called the Wigner transform. The inverse WWT is also called the Weyl transform.
31
LECTURE 3. ASYMPTOTIC EXPANSION OF DISPERSION OPERATORS 32
As a corollary (Problem 3.2), the Hermitian and anti-Hermitian parts [Eq. (C.3)] of M̂ and M satisfy
Exercise 3.2: Prove Eqs. (3.18) and (3.19). Show that the symbol of a Hermitian operator is a
Hermitian matrix or, in case of a scalar operator, a real function.
The WWT is a natural mapping between operators and functions on phase space in the following
sense. The symbol of a unit operator is unity:
W [1̂] = d s hx + s/2|1̂|x − s/2i e −i k·s
= d s hx + s/2|x − s/2i e −i k·s
= d s δ(s) e −i k·s
= 1. (3.20)
The symbol of an operator that can be represented as a function of the coordinate operator x̂ is the
same function of x:
W [F (x̂)] = d s hx + s/2|F (x̂)|x − s/2i e −i k·s
= d s F (x − s/2) hx + s/2|x − s/2i e −i k·s
= d s F (x − s/2)δ(s) e −i k·s
= F (x). (3.21)
Similarly, the symbol of an operator that can be represented as a function of the wavevector operator
k̂ is the same function of k (Exercise 3.3):
W [G(k̂)] = d s hx + s/2|G(k̂)|x − s/2i e −i k·s = G(k). (3.22)
Also, the symbol of an operator representable as F (x̂) + G(k̂) is F (x) + G(k). In summary then,
Exercise 3.3: Prove Eq. (3.22) using the result from Exercise 3.1.
In the general case, though, the correspondence ⇔ is more complicated than swapping (x̂, k̂) and
(x, k). For example,
W [k̂F (x̂)] = d s e −i k·s hx + s/2|k̂F (x̂)|x − s/2i
= d k0 d s e −i k·s hx + s/2|k0 i k̂F (x − s/2) hk0 |x − s/2i
32
LECTURE 3. ASYMPTOTIC EXPANSION OF DISPERSION OPERATORS 33
= d k0 d s e −i k·s hx + s/2|k0 i k0 F (x − s/2) hk0 |x − s/2i
=
d k0 d s k0 F (x − s/2) e i (k −k)·s 0
(2π)n
=
d k0 d s (k0 + k)F (x − s/2) e i k ·s 0
(2π)n
=k
d k0 d s F (x − s/2) e i k ·s − i d k0 d s F (x − s/2) ∂ e i k ·s
0 0
s
(2π)n (2π)n
= k d s F (x − s/2)δ(s) + i
d k0 d s ∂ [F (x − s/2)] e i k ·s 0
s
(2π)n
= kF (x) −
i ∂x
d k0 d s F (x − s/2) e i k ·s 0
2 (2π)n
= kF (x) −
i ∂x F (x). (3.24)
2
Using Eq. (3.18), one also obtains
†
W [F (x̂)k̂] = W [(F (x̂)k̂)† ]
†
= W [k̂F † (x̂)]
†
= kF † (x) − i/2 ∂x F † (x)
and as a corollary,
The terms ∂α F in Eqs. (3.24) and (3.25) emerge because x̂ and k̂ do not commute,
These terms are of order ε and vanish in the GO limit (ε → 0). Similarly, for a general M̂ = M (0) (x̂, k̂),
where M (0) is any combination of x̂ and k̂, one has
The operator D̂ serves as the envelope dispersion operator and can also be expressed in the following
invariant form:
33
LECTURE 3. ASYMPTOTIC EXPANSION OF DISPERSION OPERATORS 34
= d s hx + s/2|D̂|x − s/2i e −i [θ(x+s/2)−θ(x−s/2)+k·s] . (3.31)
sα ∂θ sα sβ ∂ 2 θ sα sβ sγ ∂3θ
θ(x ± s/2) = θ(x) ± α
+ α β
± + ..., (3.32)
2 ∂x 8 ∂x ∂x 48 ∂x ∂xβ ∂xγ
α
1 ∂ 2 k̄γ ∂3
= 1−
24 ∂xα ∂xβ ∂kα ∂kβ ∂kγ
+ . . . e −i [k̄(x)+k]·s , (3.33)
where we assumed that the second term in the parentheses is small compared to one. (A more explicit
estimate of the error is presented below.) This leads to
1 ∂ 2 k̄γ ∂3
D(x, k) = 1 −
24 ∂xα ∂xβ ∂kα ∂kβ ∂kγ
+ ... d s hx + s/2|D̂|x − s/2i e −i [k̄(x)+k]·s
1 ∂ 2 k̄γ ∂3
= 1− + . . . D x, k̄(x) + k
24 ∂xα ∂xβ ∂kα ∂kβ ∂kγ
= D 0 x, k̄(x) + k ,
(3.34)
. 1 ∂ 2 k̄γ ∂ 3 D(x, k)
D 0 (x, k) = D(x, k) − + ... (3.35)
24 ∂xα ∂xβ ∂kα ∂kβ ∂kγ
Note that the inverse WWT will, loosely speaking, turn the coordinate k into the operator k̂. The
latter will act on the wave envelope, which is considered slow in the coordinate representation, so
k̂ |Ψi = O(ε). In this sense, k = O(ε) in Eq. (3.34). Then, D 0 (x, k̄(x) + k) can be Taylor-expanded
in k, which corresponds to expanding D̂ in ε. Next, notice that
where we have assumed ∂/∂xα ∼ 1/Lc and ∂/∂ k̄ ∼ 1/k̄.5 Because we are interested only in O(ε)
corrections, this means that on the right-hand side of Eq. (3.35), the second term can be neglected
compared to the first term. In other words, D 0 (x, k) ≈ D(x, k) and thus
D(x, k) ≈ D x, k̄(x) + k , (3.37)
5A more accurate estimate is as follows. Let us consider D(x, k) in the form (3.16), i.e., D(x, k) = d s D̄(x, s) e −i k·s .
Then, ∂D/∂kα = −i d s D̄(x, s)sα e −i k·s ∼ sα D, where sα is the characteristic scale on which D̄(x, s) fades away along
the sα axis. Then, ε0 ∼ (k̄Lc )−2 (k̄s)3 ∼ ε2 (k̄s)3 . To ensure that ε0 ε, one must require (k̄s)3 ε 1. Strictly
speaking, this is an additional requirement of GO, independent of Eq. (3.5). The scale s can be estimated in the
homogeneous-plasma limit, when D(x, k) is just the dispersion tensor, D(x, k) → D(k).
34
LECTURE 3. ASYMPTOTIC EXPANSION OF DISPERSION OPERATORS 35
In practice, waves often propagate as narrow beams whose transverse scale L⊥ is much less than
the longitudinal scale Lk , which is comparable to that of the medium. For diffraction-limited
. .
beams, one has εk ∼ ε2⊥ , where εk = (k̄Lk )−1 and ε⊥ = (k̄L⊥ )−1 , so when retaining terms of the
first order in εk , one must retain effects of the second order in ε⊥ . Because ε0 [as in Eq. (3.36)] is
of order ε2k , Eq. (3.37) is still applicable, but Eq. (3.38) must be replaced with the second order
expansion in k⊥ . The corresponding envelope equation is
D x, k̄(x) Ψ −
i (∂α V α )Ψ − i V α ∂α Ψ −
1
(Φαβ : ∇⊥α ∇⊥β )Ψ = 0,
2 2
. 2
where Φαβ = ∂ D/∂kα ∂kβ evaluated at [x, k̄(x)]. This model is called the paraxial approximation
if the beam axis is a straight line, or in the general case, the quasioptical approximation [17]. It is
widely used, for example, in optics and also for modeling mm- and cm-wave beams in magnetically
confined plasmas.
where the notation is the same as in Sec. 3.3. Deviations from CCS models due to thermal effects and
ˆ to the dispersion operator D̂. Then, the symbol of
collisions often result in only small corrections ∆
this operator, D, satisfies
(0)
D − D (0) = D CCS + ∆ − D CCS + ∆(0) = ∆ − ∆(0) = O(ε∆),
(3.43)
35
LECTURE 3. ASYMPTOTIC EXPANSION OF DISPERSION OPERATORS 36
Although GO can be constructed based on a transform other than the WWT, the approxima-
tion (3.45) relies on the WWT property (3.19) that the other transform may not have. Then,
one might not be able to approximate the corresponding symbol of D̂ with the (well-known)
dispersion tensor of the wave in a homogeneous medium.
(e.g., see Ref. [21]). To the extent that the commutator [x̂, k̂] is negligible, one has W¯ [M̂ ] ≈
M (0) (x, k), so W¯ satisfies Eq. (3.28) just like the WWT. However, symbols produced by the
transform W¯ do not have the property (3.18). As a result,
(0) (0)
(W¯ [D̂])A = D A + [O(εD (0) )]A = D A + O(ε).
Both terms on the right-hand side are of the same order, so the anti-Hermitian part of the symbol
(0)
differs from the corresponding matrix in homogeneous plasma by ∼ ∂ 2 D H /∂xα ∂kα . This leads
to alternative GO equations, which often results in confusion and stirred a controversy [22, 23].
where we have used Eq. (3.28) to obtain the last equality. Because both ε and ∆ are small, the term
O(ε∆) is often negligible [as opposed to O(ε) and O(∆)], i.e., one can adopt
D ≈ D (0) . (3.44)
This means that the symbol of the dispersion operator can be approximated with the dispersion tensor
of homogeneous plasma. In fact, the two are often not even distinguished explicitly in literature.
Exercise 3.4: Calculate the symbols of: (a) D̂ E for waves in CCS nonmagnetized plasma
(Lecture 2) and (b) D̂ ψ for waves in CCS magnetized plasma (Problem PI.2).
(0)
Because D CCS is Hermitian, Eq. (3.44) is often further reduced to (Box 3.2)
(0) (0) (0) (0) (0)
D H ≈ D CCS + ∆H ≈ D CCS , D A ≈ 0 + ∆A = ∆A . (3.45)
This is justified in the following sense. In D H , which determines wave propagation (Lecture 5), the
main contribution is provided by D CCS , so ∆H is only a small correction, which is often negligible.
In contrast, D A , which determines wave dissipation (Lecture 5), is entirely determined by ∆A , so
(0)
unlike ∆H , the function ∆A must be kept. The advantage of the model (3.45) is that D CCS (Part III)
(0)
is simpler than D H (Part IV) and thus easier to implement numerically. (However, note that this
model can be insufficient when dissipation is strongly inhomogeneous. Sometimes this happens for
resonant absorption, which is naturally localized in space.)
The arguments presented in this section do not apply to corrections ∆ ˆ m caused by plasma motion.
ˆ (0) (0)
In this case, ∆m = 0, so Eq. (3.43) becomes D − D = ∆m = O(). Ignoring this correction can
lead to significant violation of energy conservation on times t = O(−1 ) and is typically not acceptable.
One must actually recalculate the dispersion operator for moving plasma to fix this issue. (But if one
needs an equation just for the energy, the problem can be bypassed as discussed in the next lectures.)
This asymmetry in how GO tolerates various corrections is due to the fact that plasma temporal
dispersion is typically much stronger than plasma spatial dispersion.
36
Lecture 4
In this lecture, we derive a complete set of equations that describe scalar waves: the local dispersion
relation, the polarization equation, the consistency relations, the ray equations, and scalar-amplitude
equations, including an equation for the wave action. These are known as the equations of GO.
Then, since ∂α = O(ε) and D A = O(ε), the envelope equation (3.41) can be simplified as follows:
D H x, k̄(x) Ψ + i D A x, k̄(x) Ψ −
i
H )Ψ − i V H ∂α Ψ = 0,
(∂α V α α
(4.2)
2
where V α α
H is the Hermitian part of V . It is readily seen from here that
This means that the field polarization must approximately satisfy the same equation as in the cor-
responding homogeneous medium, D H [x, k̄(x)]Ψ ≈ 0. Hence, it is convenient to represent Ψ in the
basis of the orthonormal eigenvectors hv (x) of D H [x, k̄(x)],
N
X
Ψ(x) = hv (x)Ψv (x), hu · hv ≡ h†u (x)hv (x) = δuv (4.4)
v=1
.
(as earlier, N = dim Ψ), and the corresponding eigenvalues Λv (x), which satisfy
37
LECTURE 4. EQUATIONS OF GEOMETRICAL OPTICS 38
Equation (4.6) shows that Ψu can be order-one only if the corresponding eigenvalue is small,
Λu = O(ε). This can be satisfied for more than one mode at a time, in which case “mode conversion”
is possible (Problem PII.2). Here, we will consider a simpler case, when only one, say uth, mode
has a noticeable amplitude. Specifically, suppose Λu = O(ε), and Λv6=u = O(1); then, Ψu = O(1)
is allowed, and Ψv6=u = O(ε). Let us re-examine the wave equation under these conditions more
carefully. Consider the product of Eq. (4.2) with h†u :
N
i
Λv h†u hv Ψv + i h†u D A hv Ψv − −i
X
0= h†u (∂α V α
H )hv Ψv h†u V α
H ∂α (hv Ψv )
v=1
2
N
(Λv δuv Ψv ) + i (h†u D A hu )Ψu −
i
H )hu Ψu − i hu V H ∂α (hu Ψu )
X
≈ h†u (∂α V α † α
v=1
2
= Λu Ψu + i Γu Ψu −
i
H )hu Ψu − i (hu V H ∂α hu )Ψu − i Vu ∂α Ψu .
h†u (∂α V α † α α
(4.7)
2
(Although u is a repeating index, no summation over u is assumed.) Here,
.
Qu =
i h
h†u V α (∂ h ) − (∂ h†
)V α
h
i
= Im
h
(∂ h†
)V α
i
H α u α u H u α u h
H u . (4.10)
2
Then, we obtain the following scalar equation:
i
Λu − Qu + i Γu − (∂α Vuα ) − i Vuα ∂α Ψu = 0. (4.11)
2
Recall now that k̄ was introduced as the gradient of θ [Eq. (3.3)] and θ has not been defined yet.
Let us define it now by requiring that2
which can be understood as a Hamilton–Jacobi equation for θ [12]. This leads to the following:
1 If N = 1, then there is only one normalized polarization vector, hu ≡ 1. Accordingly, ∂α h†u = 0, so Qu = 0.
2 This is the most common approach to formulating GO. Another natural way to define θ is to require that Ψ be
real; then Eq. (4.12) is replaced with Λu − Qu = 0. This is further discussed in Sec. 4.4.
38
LECTURE 4. EQUATIONS OF GEOMETRICAL OPTICS 39
• Equation (4.12) connects the local frequency w̄(t, x) with the local wavevector k̄(t, x). This
means that w̄ can be expressed through k̄:
D H hu = 0, (4.16)
which determines the mode polarization. This equation is not in violation of the assumed order-
ing (4.1), because the latter characterizes D H (x, k) at generic (x, k) while Eq. (4.16) characterizes
h specifically on solutions of Eq. (4.12).
• By combining Eqs. (4.11) and (4.12), one obtains the following equation for Ψu (Box 4.1):
Let us summarize these results and simplify the notation. Specifically, we now omit the bar in k̄
and drop the mode index u . Then, our main equations can be written as follows (Exercise 4.1):
This is a complete set of GO equations that allows one to find the whole vector field ψ = hΨe i θ +O(ε)
up to the O(ε) term, which is often considered negligible and otherwise can be found perturbatively.
The effect of other modes is completely ignored in this approximation, so effectively, any wave is
modeled as a scalar wave. (A more general model, which allows for linear resonant coupling of
multiple GO modes, is addressed in Problem PII.2.) Below, we discuss how to solve these equations
in practice.
39
LECTURE 4. EQUATIONS OF GEOMETRICAL OPTICS 40
If a wave beam is narrow enough such that variations of Vuα across the beam can be ignored,
then Vuα ≈ Vuα (l), Vuα ∂α Ψu ≈ Vul ∂l Ψu , and ∂α Vuα ≈ dl Vul , where l is the coordinate along the
.
beam axis. In this case, Eq. (4.17) yields the following equation for Ψ̄u = Ψu (Vul )1/2 :
This equation does not contain ∂α Vuα , so may be easier to implement numerically than Eq. (4.17).
Also notably, when Vul turns to zero, the field is singular. Beyond the GO approximation, this
corresponds to a caustic like those discussed in Sec. 2.3.
Exercise 4.1: Show that at Γ = 0, Eq. (4.21c) can be derived, along with Eq. (4.21a), from
the least-action principle with the action integral S = d x Λ(x, ∂θ)A. Here, θ and A = |Ψ|2 are
.
considered as independent functions of x and Λ is a given function.
The next step is to calculate k(t, x). This can be done by revisiting the definitions of w and k [see
Eq. (3.3), where the bars have been dropped to simplify the notation]:
. .
w(t, x) = −∂t θ(t, x), k(t, x) = ∇θ(t, x). (4.23)
The first consistency relation yields ∂kb /∂xa = ∂ka /∂xb . With this and Eq. (4.13), the second
consistency relation yields the following nonlinear PDE for k(t, x):
∂ka ∂w
=− a
∂t ∂x
∂ω ∂ω ∂kb
=− a −
∂x ∂kb ∂xa
∂ω ∂ω ∂ka
=− a −
∂x ∂kb ∂xb
In a one-dimensional system, the first consistency relation can be written as a continuity equation,
∂t k+∂x (vp k) = 0, where vp = ω/k is the phase velocity. This can be understood as a conservation
of wave crests (or zeros) in GO [24].
40
LECTURE 4. EQUATIONS OF GEOMETRICAL OPTICS 41
By taking the curl of ∂t k = −∇w, one obtains ∂t (∇ × k) = 0. Therefore, the second consistency
relation can be considered as the initial condition for the first consistency relation, much like
magnetic Gauss’s law serves as the initial condition for Faraday’s law (Lecture 1.1.1).
∂ω ∂ka
=− − vg · , (4.25)
∂xa ∂x
where the spatial derivative of ω on the right-hand side is taken at fixed k. The same can be written
in the following vector form:
In other words, the evolution of the field k(t, x) in the frame moving at the group velocity v g is
determined by the explicit dependence of ω(t, x, k) on x.
which is an ordinary differential equation (ODE) for k(t). Also notice that the instantaneous frequency
.
ω(t) = ω[t, x(t), k(t)] is governed by
dt ω = ∂t ω + ∂x ω · dt x + ∂k ω · dt k = ∂t ω + ∂x ω · ∂k ω − ∂k ω · ∂x ω = ∂t ω. (4.29)
In summary then,
Notably, Eqs. (4.30) remain applicable near reflection points as well, even though the amplitude
equation breaks down there (Box 4.1) and the GO parameter is not small. The reason for this is that
the GO approximation can be reinstated near cutoffs using phase-space rotation (Box 2.1) that does
not affect the ray equations.
. .
To better understand Eqs. (4.30), notice the following. The quantities p = ~k and H = ~ω are
commonly interpreted as the photon canonical momentum and energy. Assuming this interpretation,
the ray equations can be viewed as Hamilton’s equations of the photon motion in phase space (x, p),
with H(t, x, p) serving as the Hamiltonian:3
d x = ∂ω = ∂(~ω) = ∂H , (4.31a)
d t ∂k ∂(~k) ∂p
d p = d (~k) = − ∂(~ω) = − ∂H , (4.31b)
dt dt ∂x ∂x
d H = ~ d ω = ~ ∂ω = ∂H . (4.31c)
3 Strictly
dt dt ∂t ∂t
speaking, photons are defined through global modes and thus cannot be assigned specific coordinates. For
example, in a stationary GO wave, each photon is spread out along the whole propagation distance. However, this does
not limit the applicability of the ray equations, because they are derived from classical considerations and independently
from the photon analogy.
41
LECTURE 4. EQUATIONS OF GEOMETRICAL OPTICS 42
Equations (4.34) can also be derived as follows. Let us introduce some auxiliary time τ and
consider solutions of the original equation D̂ψ = 0 as “stationary” (∂τ ψ = 0) solutions of
(D̂ − i ∂τ )ψ = 0. The latter can be considered as D̂ ext ψ = 0, where D̂ ext = D̂ − $̂ is an
.
“extended” dispersion operator and $̂ = i ∂τ is the frequency operator associated with the new
.
time τ . Then, one can build GO for D̂ ext , in which case x0 ≡ ct and k0 ≡ −ω/c are treated
on the same footing as x and k. The corresponding dispersion relation is readily found to be
$ = Λ(t, x, ω, k), and Eqs. (4.34) emerge as the covariant ray equations in spacetime:
d xα =
∂Λ(x, k)
,
d kα =−
∂Λ(x, k)
.
dτ ∂kα dτ ∂xα
Let us define an auxiliary time variable τ via dt τ = −1/∂ω Λ.4 Using dt = (dt τ )dτ and Eqs. (4.33),
one can rewrite Eqs. (4.30) as follows:
dt ∂Λ(t, x, ω, k) d ω = + ∂Λ(t, x, ω, k) ,
dτ = − ∂ω
,
dτ ∂t
(4.34a)
d x = + ∂Λ(t, x, ω, k) , d k = − ∂Λ(t, x, ω, k) , (4.34b)
dτ ∂k dτ ∂x
where the former equation is the definition of τ included for completeness. (An alternative path to
these equations is presented in Box 4.4.) As a corollary, k is conserved if the medium is spatially
homogeneous (∂x Λ = 0), and ω is conserved if the medium is stationary (∂t Λ = 0).
∂D H
V α = h† h
∂kα
†
∂ ∂h ∂h
= (h† D H h) − D H h − h† D H
∂kα ∂kα ∂kα
†
∂Λ ∂h ∂h
= − ×0−0×
∂kα ∂kα ∂kα
4 Other definitions of τ can be used as well and lead to different but equivalent ray equations.
42
LECTURE 4. EQUATIONS OF GEOMETRICAL OPTICS 43
∂Λ
= , (4.35)
∂kα
where we have used Eq. (4.21b) and its conjugate. Then, we obtain
∂ ∂Λ 2 ∂ ∂Λ 2
|Ψ| + · |Ψ| = 2Γ |Ψ|2 , (4.36)
∂x0 ∂k0 ∂x ∂k
where the first term can as well be expressed as follows:
∂ ∂Λ 2 ∂ ∂Λ 2 ∂ ∂Λ 2
|Ψ| = |Ψ| = − |Ψ| . (4.37)
∂x0 ∂k0 ∂(ct) ∂(−ω/c) ∂t ∂ω
Let us introduce
. ∂Λ 2 ∂D H ∂D H
I= |Ψ| = h† h|Ψ|2 = Ψ† Ψ, (4.38)
∂ω ∂ω ∂ω
. ∂Λ ∂D H ∂D H
Ja = − |Ψ|2 = −h† h|Ψ|2 = −Ψ† Ψ. (4.39)
∂ka ∂ka ∂k
Importantly, one can also express J through the eigenvalue Λ as
∂k Λ
J =− I = v g I, (4.40)
∂ω Λ
where we have used Eq. (4.33). Then, Eq. (4.36) can be expressed as a continuity equation for I,
∂t I + ∇ · (v g I) = 2γI, (4.41)
where γ is given by
. Γ h† D A h h† D A h
γ=− =− † =− . (4.42)
∂ω Λ h (∂ω D H ) h ∂ω Λ
The quantity I that enters this equation [and is given by Eq. (4.38)] is called the action density,
and accordingly, J can be understood as the action flux density. The physical meaning of these
quantities will be discussed in detail in Lecture 5. Meanwhile, note that Eq. (4.41) can be rewritten as
[∂t + (v g · ∇)] I = −I∇ · v g + 2γI. (4.43)
.
This readily leads to an ODE for the action density on a ray, I(t) = I[t, x(t)], namely,5
dt I = 2γeff (t)I, .
γeff (t) = (γ − ∇ · v g /2)[t,x(t),k(t)] . (4.44)
By integrating Eq. (4.44), one obtains
t
I(t) = exp d t 2γeff (t ) I0 ,
0 0
(4.45)
0
43
LECTURE 4. EQUATIONS OF GEOMETRICAL OPTICS 44
∗
4.4 Spin Hall effect of light
The dispersion relation (4.46) is postulated in a somewhat arbitrary manner and is not entirely
consistent. The term i Q in the amplitude equation (4.17) can, with enough time, produce an arbitrarily
large gradient of arg Ψ. Because Q does not enter Eq. (4.41), such a gradient has no effect on |Ψ| within
the GO approximation, However, it can eventually undermine the validity of the GO approximation,
because the latter requires that Ψ be a slow function.
To prevent this, one can define θ such that Ψ be real. Then, by taking the real part of Eq. (4.11),
one arrives at a dispersion relation
instead of Eq. (4.12). This complicates the polarization equation, which now becomes
D H h = Λh = Qh (4.48)
[cf. Eq. (4.35); here we have substituted Eq. (4.48) and h† h = 1]. However, remember that redefining
θ entails modification of k, so V α is now evaluated on a different k and thus is affected. Thus in
reality, a wave propagates somewhat differently than as predicted by the conventional ray equations
(4.30). This effect is known as the spin Hall effect of light and has been observed experimentally [25].
The spin Hall effect is typically negligible in practical applications, so it is usually ignored or not
even recognized. Still, this effect is interesting because of its analogy with the spin–orbital interaction
known for quantum particles. (The GO limit in quantum mechanics is known as the semiclassical
approximation, and ray equations for quantum waves are known as classical mechanics.) When placed
in an external magnetic field, quantum particles exhibit different trajectories depending on their spin
state, for example, as seen in the famous Stern–Gerlach experiment. The analog of the spin for
electromagnetic waves is their polarization state, and parameters of the medium serve as the vector
potential, whose derivatives create an effective “magnetic field” for the “spin” to interact with. This
analogy can be made quantitative and in fact the spin Hall effect of light and the Stern–Gerlach effect
are mathematically identical [20, 26, 27].
44
Lecture 5
In this lecture, we discuss the physical meaning of the wave action and also introduce the wave energy,
the wave momentum, and the corresponding transport equations.
∂t I + ∇ · (v g I) = 2γI, (5.1)
where I = Ψ† (∂ω D H )Ψ is the wave action density and γ = −h† D A h/∂ω Λ. By integrating Eq. (5.1)
over the whole space, one obtains
d d x I + d x ∇ · (v I) = d x 2γI. (5.2)
dt g
By Gauss’s theorem, the second term equals the flux through the infinite surface and thus is zero,
.
assuming that the field is localized. Then, if γ = 0, the total action I = d x I is conserved:
dI = 0. (5.3)
dt
The reason for this conservation law is that at zero γ, the wave is a Lagrangian system whose La-
grangian density does not depend on θ explicitly (Box 5.1). This makes I a Noether invariant of an
asymptotic theory, an adiabatic invariant. The well-known adiabatic invariant of a harmonic oscillator
can be understood as a special case of I (Box 5.2). In a broader context, Eqs. (5.1)–(5.3) generalize the
WKB conservation law (2.35) that was derived in Lecture 2 for a specific wave in a simple geometry.
Because I satisfies a continuity equation (modulo the local dissipation term γ), it can be interpreted
as the density of quasiparticles that travel with velocity v g . With the same reservations as in Lecture 4,
these particles can be identified as photons. We will return to the photon analogy later in this lecture.
45
LECTURE 5. WAVE ACTION, ENERGY, AND MOMENTUM 46
where the mode index is omitted and Γ ∝ γ has vanished. The operator Lˆ is Hermitian, so this
equation satisfies the least-action principle δS[Ψ , Ψ] = 0 with S = d t d x Ψ∗ LˆΨ, where Ψ and
∗
which are the anticipated GO dispersion relation and the action-conservation theorem. Note that
there is no need to retain O(ε) corrections when deriving the action-conservation theorem from
.
a variational
principle. Also note that one obtains a coninuity equation for I = ∂ω L for any
S ≈ d t d x L such that the wave Lagrangian density has the form L = L(x, ∂θ). However, if L
explicitly depends on θ rather than ∂θ, then the wave action is not conserved.
a Because Ψ and Ψ∗ are linearly independent combinations of Re Ψ and Im Ψ, and because Re Ψ and Im Ψ can
be treated as independent, the complex functions Ψ and Ψ∗ can be treated as independent too.
where is the symbol of ˆ. Also, from Faraday’s law the magnetic-field complex envelope is, to the
leading order, B = (c/ω)(k × E ).
The corresponding action density can be calculated as follows:
2
2c2
1 † c † 2 1 † † 2
I= E ∂ω 2 (kk − 1 k ) + H E = E − 3 (kk − 1 k ) + ∂ω H E . (5.10)
16π ω 16π ω
factor (16π)−1 (compared to Lecture 1) in order to simplify the interpretation of I in later sections.
46
LECTURE 5. WAVE ACTION, ENERGY, AND MOMENTUM 47
Our formulation of GO applies not just to electromagnetic waves but also to any linear oscil-
lating systems that satisfy the assumed orderings. For example, consider a dissipative harmonic
oscillator with time-dependent frequency ω0 (t):
q̈ + 2ν q̇ + ω02 (t)q = 0, q = Re (a e i θ ),
assuming the damping coefficient ν is a small constant. We can rewrite this as D̂q = 0, where
D̂ = ω̂ 2 + 2i ν ω̂ − ω02 (t̂), ω̂ = i ∂t .
.
Since D is a scalar, one has Λ = DH = ω 2 − ω02 (t), so the local dispersion relation is ω 2 = ω02 (t).
The action density I, which is the same as the total action I here, is given by
.
I = (∂ω Λ) |a|2 = 2ω0 |a|2
1
= E † (ω∂ω H + 2H )E
16πω
1
= E † ∂ω (ω 2 H )E . (5.11)
16πω 2
Using that
1 h c2 i
I= E † − 2 (kk† − 1 k2 ) − (DH − H ) + ω∂ω H E
16πω ω
1 †
h c2 i
= E ∂ω (ωH ) − 2 (kk† − 1 k2 ) E
16πω ω
1 h † c2 i
= E ∂ω (ωH )E + 2 |k × E |2
16πω ω
1 h † i
= E ∂ω (ωH )E + |B |2 . (5.12)
16πω
47
LECTURE 5. WAVE ACTION, ENERGY, AND MOMENTUM 48
Then in summary,
1 1
I= 2
E † ∂ω (ω 2 H )E = [E † ∂ω (ωH )E + B †B ], (5.13)
16πω 16πω
Exercise 5.1: Consider any a = Re (ac e i θ ) and b = Re (bc e i θ ) with slow complex envelopes ac
and bc and fast real phase θ. Show that their θ-averaged product can be expressed as follows:
. c2 ∂
Sg = − (ka kb − δab kc kc )E ∗a E b
16πω ∂kg
c2
=− (δag kb + ka δbg − 2δab δcg kc )E ∗a E b
16πω
c2
= (2kg E ∗a E a − E ∗g kb E b − E g ka E ∗a )
16πω
c2
= [E × (k × E ∗ ) + E ∗ × (k × E )]g
16πω
c2
= Re [E × (k × E ∗ )]g
8πω
c
= Re (E × B ∗ )g , (5.16)
8π
so S is the average Poynting vector,
c c
S= Re (E × B ∗ ) = hẼ × B̃i. (5.17)
8π 4π
Then K, whose components are
. ω ∂
Ka = − E† HE, (5.18)
16π ∂ka
must be the kinetic action flux density. We will symbolically express it in the vector form as follows:
ω ω †
K=− E † ∂k HE = − hẼ ∂k H Ẽi. (5.19)
16π 8π
48
LECTURE 5. WAVE ACTION, ENERGY, AND MOMENTUM 49
49
LECTURE 5. WAVE ACTION, ENERGY, AND MOMENTUM 50
U = ~ωnph , (5.29)
where nph is the photon density. It is seen from here that I = ~nph ; i.e., I is just the photon density
in units ~−1 (cf. Exercise 5.2). Accordingly, Eq. (5.21) can be understood as follows. Let us assume
for now that γ = 0. Then, the action conservation (5.1) says that the number of photons (action/~) is
conserved, and thus the wave energy can evolve only through the evolution of the energies of individual
photons, H = ~ω. This regime is called adiabatic (cf. Box 5.2). The corresponding power density is
Exercise 5.2: One can use the Taylor expansion of Eq. (5.6) around the solution of the
dispersion relation w = ω(t, x, k) to obtain Λ(t, x, w, k) ≈ (w − ω(t, x, k))∂ω Λ. Then, using
∂ω Λ|Ψ|2 = I, one arrives at the “canonical” form of the wave Lagrangian density [24, 30]
where the independent functions are θ and I. Show that in the “point-particle limit”, when
I(t, x) = ~δ[x − X(t)],a this action becomes S[X, P ] = d t [P · Ẋ − H(t, X, P )] and leads to
Hamilton’s equations (4.31), with H = ~ω(t, X, P ) as the Hamiltonian, X(t) as the coordinate,
.
and P (t) = ∇θ[t, X(t)] as the canonical momentum [31].
a Here, “δ[x − X(t)]” denotes a profile that is narrow compared to the scale of the background medium but
Dissipative waves (γ 6= 0)
Now, let us allow for non-conservation of photons, i.e., nonzero γ. To the extent that U still can be
.
understood as the wave energy density,3 the function Pabs = −2γ U can be interpreted as the wave
power density absorbed irreversibly. (The sign is chosen such that Pabs > 0 when a wave is losing
action.) This is understood because dissipation is due to the loss of photons; the loss of a single
photon results in the energy loss equal to ~ω, so the loss of 2γI/~ photons per unit time per unit
volume results in the loss of energy 2γ U .
Using Eq. (4.42), one finds that
h† A h h† A h
1 ω ω
γU = − ωI = − ∂ω Λ |E |2 = − E † AE . (5.32)
16π ∂ω Λ 16π ∂ω Λ 16π
Hence,
ω †
Pabs = E AE . (5.33)
8π
3 Strictly speaking, the concept of energy is undefined for dissipative systems. However, if such a system transitions
from a conservative state into another conservative one, then the change of its energy is well defined. The function
U can be used as a means to calculate this change, because it represents the true energy density before and after the
transition and its governing equation (5.21) holds at all times. The physical meaning of U during the transition is
irrelevant, but one might as well call U energy density for the lack of a better definition of the energy density.
50
LECTURE 5. WAVE ACTION, ENERGY, AND MOMENTUM 51
where j̃ ind and Ẽ = Re (E e i θ ) are the real current density and the real electric field (Exercise 5.1).
Therefore, Pabs represents the Joule-heating power.
∂t P a + ∇ · (P a v g ) = −I∂a ω + 2γ P a . (5.37)
Suppose there is no dissipation (γ = 0), in which case the system becomes Lagrangian. Also suppose
that the medium is homogeneous (∂a ω = 0). Then, this equation becomes conservative:
∂t P a + ∇ · (P a v g ) = 0, d x P a = const. (5.38)
In other words, theinvariance of a Lagrangian system with respect to translations in time leads to
the conservation of d x P . This means that by definition [12], P is the density of the wave canonical
energy, at least up to a constant factor. This also means that the energy of a linear wave propagates
at the group velocity. One can also extend this discussion by analogy with Sec. 5.2.
Finally, notice that together, Eqs. (5.36) with Eq. (5.22) lead to the following fundamental relation
between the wave momentum density and the wave energy density:
k
P= U. (5.39)
ω
Equation (5.39) is similar to the relation between photon momentum p = ~k and the photon energy
H = ~ω. Because of this, some authors derive Eq. (5.39) from the quantum analogy. However, as
shown above, Eq. (5.39) can be derived using purely classical arguments as well.
51
LECTURE 5. WAVE ACTION, ENERGY, AND MOMENTUM 52
k k
A
B
z
r k
Figure 5.1: Schematic of the geometry assumed in Sec. 5.4.
assume the planar geometry but adopt the notation for the coordinates as in the cylindrical geometry
such as that in a tokamak with a high aspect ratio. Assume that the vector potential A has the form
A = eθ Ψ(r), so B 0 = ez Ψ0 (r) (Fig. 5.1). A radial static electric field is also allowed but will not be
important. Suppose that a wave has deposited energy ∆U through the interaction with a particle with
charge q. Since the overall system is symmetric in θ, the total canonical momentum in the θ direction
must be conserved. Hence, the particle must change its canonical momentum in the θ direction by
k⊥ ∆U/ω. This leads to
h q i k
⊥
∆ pθ + Ψ(r) = ∆U, (5.40)
c ω
where p is the particle kinetic momentum. After the interaction, the left-hand side does not change,
so it is equal to its own time average. But hpθ i = 0 (the average perpendicular velocity of a particle
that experiences stationary Larmor rotation in a dc magnetic field is zero), so one obtains
52
Problems for Part II
(a) Write down the ray equations explicitly. Find the wave frequency ω(x), the wavevector k(x),
and the group velocity vg (x). At what x is the wave reflected?
(b) Find and plot the ray trajectory x(t). Find the time at which the envelope arrives at the
reflection point.
(c) Calculate the wave energy density. How much of it is stored in: (i) the electric field, (ii) the
magnetic field, and (iii) plasma oscillations?
(d) Write down the conservation law for the wave energy in a differential form. Using the above
results, find the electric field amplitude as a function of x for a stationary wave. Compare the
result with Eq. (2.35) obtained from the WKB approximation.
(e) Suppose now that the plasma is homogeneous but undergoes mechanical compression trans-
versely to the wavevector. What is conserved in this case? Assuming for simplicity that ∂x ne = 0,
find the wave total energy as a function of the instantaneous density ne (t).
i
Λ − Q + i Γ − (∂α V ) − i V ∂α Ψ̄ ≈ 0,
α α
(5.44)
2
53
PROBLEMS FOR PART II 54
which you may consider as given. Here, Λ and V α are 2 × 2 real diagonal matrices given by
Λ1 0 ∂Λ1 /∂kα 0
Λ= , Vα≈ , (5.45)
0 Λ2 0 ∂Λ2 /∂kα
Q is a nondiagonal Hermitian matrix that determines mode coupling, and Γ is a (generally, also non-
diagonal) Hermitian matrix that determines dissipation. For simplicity, let us suppose that dissipation
is negligible (Γ ≈ 0) and that the problem is one-dimensional, i.e., all functions depend on only one
spacetime coordinate l (which can be the time, a spatial coordinate, or a linear combination thereof).
Suppose also that ∂Λ1 /∂kl and ∂Λ2 /∂kl have the same sign, say, are positive.
.
p p
(a) Consider ζ = V l Ψ̄, where V l is a (diagonal) matrix whose square equals V l . Show that ζ
satisfies the following equation, where H is a Hermitian 2 × 2 matrix:4
(b) Because H is Hermitian, Eq. (5.46) has a conservation law. Write down this law in terms of ζ,
then express it in terms of Ψ̄. What is the physical meaning of the conserved quantity?
(c) Like in Eq. (4.11), k̄(x) must be chosen such that the right side of Eq. (5.46) be small, namely, O(ε).
We cannot choose k̄(x) to make the whole Λ zero, because Λ is now a matrix that is determined
by two independent scalar functions, Λ1 and Λ2 . But by our assumption that ψ is quasi-
monochromatic, the two modes must be approximately in resonance; thus, if we choose k̄(x),
say, such that Λ1 be small, then Λ2 will be small automatically, and vice versa. Like in Lecture 4,
the specific choice of k̄(x) is a matter of convenience; for example, one can choose k̄(x) such that
c1 Λ1 + c2 Λ2 = 0, where c1,2 can have any values of order one or less. It is conventional to adopt
tr H[x, k̄(x)] = 0; then, H can be parametrized as follows:5
−α −i β
!
H=
i β∗ α , (5.47)
where α is real and α, β = O(ε). Show that the variable transformation a1 = ζ1 e −i γ/2 and
.
a2 = ζ2 e
. i γ/2 .
, where γ = arg β, leads to the following equation for a1,2 :
.
where τ = d l |β| and u = α/|β|− γ̇/2, and the dot denotes d /d τ . Equation (5.48) is the canon-
.
ical form of the envelope equation describing one-dimensional coupling of two (nondissipative)
resonant modes. Examples of such “mode conversion” in plasma will be discussed in Part III.
(d) For constant u, search for the eigenmodes of Eq. (5.48) in the form a1,2 = A1,2 e −i Ωτ . The
values of A1 and A2 characterize how close these eigenmodes are to Mode I and Mode II of the
corresponding homogeneous medium (Q = 0). [For example, if A = (1, 0)| , then the value of a2
in such eigenmode is zero, i.e., the eigenmode is purely Mode I.] Plot Ω and A1,2 as functions
of u for both eigenmodes and explain where each eigenmode is close to Mode I and where it is
close to Mode II.
4 Notably, Eq. (5.46) is similar to the equation governing a two-level quantum system.
5 Choosing an alternative convention leads to equations that have are equivalent to Eq. (5.48) but a different form.
This is due to the fact that redefining k̄(x) implies that the total phase of ψ is split differently between Ψ and θ,
resulting in a different definition of the envelope functions a1,2 . In other words, the alternative equations are different
because they describe the same total field ψ in different variables.
54
PROBLEMS FOR PART II 55
(e) Suppose u(τ ) = µτ , where µ is constant, and assume µ > 0 for simplicity. Here, τ = 0 is
understood as the moment when Mode I and Mode II are exactly in resonance. In quantum
mechanics, this is known as the Landau–Zener problem [32, 33]. One way to solve Eq. (5.48) in
this case is to rewrite it as the Weber equation, ä1 + (1 + µ2 τ 2 − i µ)a1 = 0, and then explore the
asymptotic behavior of its solutions, which are given by parabolic cylinder functions. But let us
adopt a more intuitive approach (which is an abridgment of the approach used in Ref. [34]).
First, show that Eq. (5.48) can be expressed as follows:
a1 . s − i ∂s . s + i ∂s . √
ξˆη̂ a1 = − , ξˆ = √ , η̂ = √ , s = τ µ. (5.49)
2µ 2 2
For clarity, let us call s a “time” variable, so i ∂s can be viewed as the s-representation of the
corresponding “frequency” operator ω̂ s . (Alternatively, one can view s as a spatial variable
and −i ∂s as the corresponding wavevector operator, or momentum √ operator.) Then, √ one can
write ξˆ and η̂ in the following invariant form: ξˆ = (ŝ − ω̂ s )/ 2 and η̂ = (ŝ + ω̂ s )/ 2. Show
ˆ η̂] = −i . This is identical to [ŝ, ω̂ s ], so one can view ξˆ and η̂ as the new time and
that [ξ,
frequency operators in the phase space rotated by 45◦ with respect to (s, ωs ).6 Accordingly, in
the ξ-representation, Eq. (5.49) has the form
i ξ ddaξ1 =−
a1
2µ
. (5.50)
(You are not asked to prove this formally but you can try to do it if you are interested; otherwise,
.
see the solutions later.) Integrate Eq. (5.50) and show that the “transmission coefficient” T =
a1 (+∞)/a1 (−∞) is given by7
T = e −π/(2µ) . (5.51)
Hint: In order to encircle the pole at ξ = 0 correctly, you will need to introduce
infinitesimal positive dissipation, i.e., replace ∓u with ∓u − i 0 [cf. the discussion
preceding Eq. (2.20)]. If you did not study complex analysis, know that for any g,
b b
lim
ν→0+
d ξ ξg(ξ)
+ iν
= −i πg(0) + d ξ g(ξ)
ξ
, (5.52)
a a
where d ξ (...) is the Cauchy principal value of the corresponding integral. This is
known as the Sokhotski–Plemelj theorem. It will be discussed in detail in Part IV.
(f) Using Eq. (5.51) and the result of part (b), qualitatively explain the result of mode conversion
at |µ| 1 (“adiabatic” regime) and at |µ| 1 (“diabatic” regime).
.
(g) Show that µ is of the order of the largest relative rate at which the beat period Tb = 2π/|ΩI −ΩII |
evolves with τ ; i.e., show that |µ| ∼ max |d Tb /d τ |. Using this, formulate the condition under
8
which the interaction between two modes can be neglected, i.e., the single-mode model studied
in Lecture 4 can be used instead of the more complicated two-mode theory discussed above.
55
Part III
In this part of the course, we overview basic waves in plasmas within fluid models. Our
overview is intended as introductory rather than exhaustive. For more information, see,
for example, Refs. [1, 3, 35].
56
Lecture 6
In this lecture, we study waves in cold magnetized plasma in the same manner as we studied waves
in cold nonmagnetized plasma in Lecture 2.
57
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 58
In case of a homogeneous plasma, one can assume all the “tilded” quantities to be ∝ e −i ωt+i k·x and
infer the spectral conductivity as described in Lecture 1. More generally, though, we are interested
in the Weyl symbol of σ̂, which is derived as follows. Equation (6.4) has the same form as Eq. (2.5),
(f) (i)
and similarly, its general solution of Eq. (6.4) can be written as j̃s,± = j̃ s,± + j̃ s,± ,
(Here and further, ωps , Ωs , and νs are evaluated at location x.) Then, from Eq. (6.5), one finds
2 t
ωps
d t0 + cos[Ωs (t − t0 )] Ẽx (t0 , x) + sin[Ωs (t − t0 )] Ẽy (t0 , x) e −νs (t−t ) ,
n o 0
j̃ (i)
s,x (t, x) = (6.7a)
4π t0
2 t
ωps
d t0 − sin[Ωs (t − t0 )] Ẽx (t0 , x) + cos[Ωs (t − t0 )] Ẽy (t0 , x) e −νs (t−t ) ,
n o 0
j̃ (i)
s,y (t, x) = (6.7b)
4π t0
The right-hand side is independent of t, which is due to the fact that the plasma is assumed stationary;
however, it generally depends on x through ωps , Ωs , and νs . After substituting the corresponding fab
and performing the integration, one obtains
σs,xx = +σs,yy = +
2
ωps i (ω + i νs ) , (6.11a)
4π (ω + i νs )2 − Ω2s
2
ωps Ωs
σs,xy = −σs,yx =− , (6.11b)
4π (ω + i νs )2 − Ω2s
58
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 59
2
ωps
σs,zz = − , (6.11c)
4π i (ω + i νs )
where the time integrals (6.10) converge because Im ω = 0 and νs > 0. Because this result extends to
homogeneous plasma as is, we will ignore the distinction between the symbol of σ̂ from the spectral
conductivity in this lecture. For the same reason, we will not distinguish the symbol of χ̂ from the
spectral susceptibility and the symbol of ˆ from the dielectric tensor. We will also assume for simplicity
that the plasma is collisionless, i.e., νs → 0+ [cf. Eq. (2.20)].
χs,xy = −χs,yx =−
i Ωs ωps 2
, (6.12b)
ω ω 2 − Ω2s
2
ωps
χs,zz =− 2. (6.12c)
ω
Using Eqs. (6.12), one readily obtains the corresponding dielectric tensor (Box 6.1):
S −iD 0
= iD S 0 , (6.13)
0 0 P
which is Hermitian.1 The notation assumed here is the traditional notation from Ref. [1]:
2
ωps
S =1− = (R + L),
1
X
(6.14a)
s
ω 2 − Ω2s 2
2
ωps
D= = (R − L),
X Ωs 1
(6.14b)
s
ω ω2 − Ω2s 2
2
X ωps
P =1− ω2
, (6.14c)
s
2
ωps
R, L = 1 −
X
. (6.14d)
s
ω(ω ± Ωs )
If ω is much larger than all Ωs and ωps , then ≈ 1. In the somewhat more general case when
ω Ωe yet ω ∼ ωpe , the tensor is the same as in nonmagnetized plasma. Let us also consider the
low-frequency limit, ω Ωi , which is known as the magnetohydrodynamic (MHD) limit. There, one
has S ≈ 1 + γA , where
2 X 4πns0 e2 m2 c2
. X ωps s s 4π X c2
γA = 2
= 2 2 = c2 2 ns0 ms = 2 . (6.16)
s
Ωs s
m s es B0 B0 s VA
59
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 60
Using the Gell–Mann matrices α introduced in Problem PI.2, one can also express in the
following invariant form that allows for an arbitrary orientation of B 0 :
" #
2 2 2
X ωps ωps (α · Ωs ) ωps (α · Ωs )2
=1+ −1 2 + − 2 2 . (6.15)
s
ω ω(ω 2 − Ω2s ) ω ω − Ω2s
. P
and the mass density ρm = s ns0 ms . Also,
2
ωps
D=−
X Ωs
s
ω Ω2s
(1 − ω 2 /Ω2s )
2
X Ωs ωps ω2
≈− 1+ 2
s
ω Ω2s Ωs
2 2
1 X ωps X ω ωps
=− − . (6.18)
ω s Ωs s
Ωs Ω2s
Here,
2
X ωps X 4πn0,s e2 ms c
s 4πc X
= = n0,s es = 0 (6.19)
s
Ωs s
ms es B0 B0 s
Hence, we obtain the following low-frequency limit of , which we will use later (Exercise 6.1):
1 + γA 0 0
≈ 0 1 + γA 0 . (6.21)
0 0 P
Exercise 6.1: At ω → 0, the wave field becomes stationary, and one might expect that a
stationary field cannot create a current perpendicular to B 0 . Nevertheless, the above calculation
shows that xx = yy → 1 + γA , which is not unity. (In fact, γA is often large.) This means that
plasma does respond to such field. What is the physical nature of this response?
60
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 61
Without loss of generality, one can choose the coordinate axes such that k = (k⊥ , 0, kk ), where
k⊥ = k sin θ, kk = k cos θ, and θ is the angle between k and B 0 . A similar notation will be assumed
for the refractive index N . Then, the dispersion tensor (or the symbol of D̂) is given by
S − Nk2 −iD N⊥ Nk
DE = iD S − N2 0 , (6.23)
N⊥ Nk 0 P − N⊥2
and thus,
det D E
= (S − N 2 cos2 θ)(S − N 2 )(P − N 2 sin2 θ) − (−iD )(iD )(P − N 2 sin2 θ) − (N 2 sin θ cos θ)2 (S − N 2 )
= (S − N 2 )(SP − S N 2 sin2 θ − P N 2 cos2 θ + N 4 sin2 θ cos2 θ − N 4 sin2 θ cos2 θ) − D 2 (P − N 2 sin2 θ)
= (S − N 2 )(SP − S N 2 sin2 θ − P N 2 cos2 θ) − D 2 (P − N 2 sin2 θ)
= S 2 P − S 2 N 2 sin2 θ − SP N 2 cos2 θ − SP N 2 + S N 4 sin2 θ + P N 4 cos2 θ − D 2 P + D 2 N 2 sin2 θ
= N 4 (S sin2 θ + P cos2 θ) + N 2 (−S 2 sin2 θ − SP cos2 θ − SP + D 2 sin2 θ) + S 2 P − D 2 P
= N 4 (S sin2 θ + P cos2 θ) − N 2 [(S 2 − D 2 ) sin2 θ + PS (1 + cos2 θ)] + P (S 2 − D 2 ).
and introduce
AN 4 − BN 2 + C = 0. (6.26)
Note that F 2 > 0 at all real frequencies, so N 2 is real too. Thus, N is either real or imaginary.
.
Consider also τ = tan2 θ and note that
1 τ
cos2 θ = , sin2 θ = . (6.28)
1+τ 1+τ
61
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 62
Then, we obtain
This leads to the following handy formula, which will be used below:
τ =−
P (N 2 − R )(N 2 − L) . (6.30)
(N 2 S − RL)(N 2 − P )
6.4 Eigenmodes
Solutions of the dispersion relation (6.27) can be qualitatively understood by analyzing the charac-
teristic frequencies and limits, which are as follows.
The former is satisfied when ω 2 = ωp2 . The other two equations generally have multiple solutions
depending on the number of species; but there are only two solutions in case of electron–ion plasma
with single type of ions, ω = ωR ,L , and ωR ωL .
Other notable frequencies are resonances, which correspond to N → ∞. Resonances can be found
by taking the corresponding limit in Eq. (6.30) or, equivalently, by noticing that our solution for N 2
predicts infinite refractive index at A = 0. In either case, one finds that such points are located where
This equation coincides with the electrostatic dispersion relation discussed in Problem PI.1. How-
ever, note that the resonance condition is necessary but insufficient for a field to be electrostatic
(Exercise 6.2).
Exercise 6.2: Show that the electrostatic dispersion relation discussed in Problem PI.1 leads
to Eq. (6.32). Using the condition (2.57), explain which resonances in cold magnetized plasma
(discussed below) are electrostatic and which are not.
62
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 63
ω ω
Ωe ωp
ωp Ωe
Ωi Ωi
θ θ
π/8 π/4 3π/8 π/2 π/8 π/4 3π/8 π/2
Figure 6.1: Resonance frequencies in electron–ion plasma with one type of ions [numerical solution of
Eq. (6.32)]. Left – underdense plasma (ωp < |Ωe |). Right – overdense plasma (ωp > |Ωe |). In both
cases, there are three resonances at θ = 0 (at frequencies ωp , Ωe , and Ωi ; all dashed), and there are
two resonances at θ = π/2 (at frequencies ωlh and ωuh ).
corresponds to oscillations at frequencies high enough for the ion response to be negligible, so ωuh can
be found from
X ωps 2 2
ωpe
0=1− ≈ 1 − . (6.33)
s
ω 2 − Ω2s ω 2 − Ω2e
In electron–ion plasma with only one type of ions, the only other hybrid resonance is the lower-hybrid
resonance that corresponds to the frequency (Exercise 6.3)
!−1/2
1 1
ωlh = 2 + Ω2 + |Ω Ω | . (6.35)
ωpi i e i
At intermediate θ, the corresponding resonances are shown in Fig. 6.1. Plasmas with multiple ion
species also have ion–ion hybrid resonances.
Exercise 6.3: Consider electron–ion plasma with single type of ions with charge ei = Ze. Show
that up to terms of order Zme /mi , S can be approximated as follows:
2 2 2 2
− ωuh )(ω − ωlh )
S ≈ (ω
(ω 2 − Ω2 )(ω 2 − Ω2 )
,
e i
where ωuh is defined in Eq. (6.34) and ωlh is defined in Eq. (6.35).
63
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 64
At small enough ω, we also have large P 3 while the refraction index remains finite, as we will find
shortly. Then, hz ≈ −(N⊥ Nk /P )hx hx , so the first two field equations can be written as follows:
1 + γA − Nk2
0 hx
= 0. (6.37)
0 1 + γA − N 2 hy
Equation (6.37) indicates that there are two modes in this limit. The first one, known as the shear
Alfvén wave, is x-polarized (hy = 0) and satisfies
Nk2 = 1 + γA , (6.38)
or equivalently,
kk2 VA2
ω2 = −1 . (6.39)
1 + γA
The second mode, known as the compressional Alfvén wave, is y-polarized (hx = 0) and satisfies
N 2 = 1 + γA , (6.40)
or equivalently,
k 2 c2 k 2 c2 k 2 VA2
ω2 = = −1 = −1 . (6.41)
1 + γA γA (1 + γA ) 1 + γA
Many plasmas of practical interest, including magnetically confined fusion plasmas, have γA 1,
−1
so the term γA in Eqs. (6.39) and (6.41) is often neglected. These waves are further studied in
Problem PIII.3.
64
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 65
N2 N2
1 + γA 1 + γA
1 1
ω ω
0 Ωi ωlh ωp Ωe ωuh 0 Ωi ωlh ωp Ωe ωuh
ω ω
ωuh ωuh
Ωe Ωe
ωp ωp
ωlh ωlh
Ωi Ωi
k k
0 0
Figure 6.2: The dispersion curves of underdense plasma: N 2 (ω) (upper row) and ω(k) (lower row).
The left column corresponds to θ = 0◦ : red – L mode, blue – R mode, green – Langmuir oscillations.
The right column corresponds to θ = 90◦ : red – X mode, blue – O mode. The dashed diagonals
correspond to ω = ck.
65
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 66
N2 N2
1 + γA 1 + γA
1 1
ω ω
0 Ωi ωlh Ωe ωp ωuh 0 Ωi ωlh Ωe ωp ωuh
ω ω
ωuh ωuh
ωp ωp
Ωe Ωe
ωlh ωlh
Ωi Ωi
k k
0 0
Figure 6.3: The dispersion curves of overdense plasma: N 2 (ω) (upper row) and ω(k) (lower row). The
left column corresponds to θ = 0◦ : red – L mode, blue – R mode, green – Langmuir oscillations. The
right column corresponds to θ = 90◦ : red – X mode, blue – O mode. The dashed diagonals correspond
to ω = ck.
66
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 67
Exercise 6.4: Which direction do the R and L waves rotate relative to the particle rotation?
Can you answer this without doing calculations?
S −iD
0 hx
iD S − N 2 0 hy = 0, (6.46)
0 0 P − N2 hz
and the equation for τ has two solutions (Figs. 6.2 and 6.3). The first one,
N 2 − P = 0, (6.47)
ω 2 = ωp2 + c2 k 2 , (6.48)
with a cutoff at P = 0. The corresponding mode is called the O wave (“ordinary wave”). The fact
that the O wave is insensitive to B0 is explained by the wave polarization. One can see from
S −iD 0
hx
iD S − N 2 0 hy = 0 (6.49)
0 0 0 hz
that the corresponding E is parallel to B 0 . Such field causes oscillations of (cold) particles parallel
to B 0 , so the magnetic Lorentz force on the particles in the O wave is zero.
The other mode, called the X wave (“extraordinary wave”), corresponds to
N 2 = RL/S . (6.50)
It has cutoffs at R = 0 and L = 0. The resonances are zeros of S, i.e., hybrid resonances. The
polarization is found from
S −iD
0 hx
iD (S 2 − RL)/S 0 hy = 0. (6.51)
0 0 P − (S − RL)/S
2
hz
This shows that hz = 0 and hx /hy = iD /S , so the X-wave polarization is elliptic in the (x, y) plane. At
the electron cyclotron resonance in particular, one has S ≈ R /2 ≈ D , so hx /hy ≈ i , which corresponds
to a circular polarization. Similarly, at ion cyclotron resonances, one has a circular polarization in the
opposite direction, hx /hy = −i .
67
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 68
N2 N2
1 + γA
1 + γA 1
ω
0 Ωi ωlh Ωe ωp ωuh
ω
0 Ωi ωlh ωp Ωe ωuh
ω
ω
ωuh
ωuh
Ωe ωp
Ωe
ωp
ωlh ωlh
Ωi Ωi
k k
0 0
Figure 6.4: The dispersion curves for various θ (colder colors correspond to smaller θ): N 2 (ω) (upper
row) and ω(k) (lower row). The left column corresponds to underdense plasma, the right column
corresponds to overdense plasma. The dashed diagonals correspond to ω = ck.
68
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 69
Exercise 6.5: Qualitatively plot the dispersion curves for plasma that has two types of ions.
∗
6.4.6 Level repulsion
In this (optional) section, we explain the cause of level repulsion as a generic effect. Consider a general
wave system governed by DΨ = 0 with Hermitian D. Suppose there are two waves that are in exact
resonance (have their dispersion curves crossed) at some ω = ω0 , k = k0 , and θ = θ0 . Let us perturb
. .
k by some small δk = k − k0 and θ by some small δθ = θ − θ0 and consider the resulting frequency
.
shifts of the eigenmodes, δω = ω − ω0 . These shifts are determined by the field equation that for
homogeneous waves can be written as follows:
Λ + Ξ† (δD) Ξ Ψ̄ = 0,
(6.53)
and the second term is small so it can be evaluated at ω = ω0 . Because V ω and V k are diagonal and
real, and because Ξ† (δD) Ξ is Hermitian, one can readily show that v1,2 and ∆1,2 are real and
where β is a complex number. The corresponding frequency shifts δω are found from det(H̄−1 δω) = 0.
The latter is a quadratic equation for δω, so it has two roots that satisfy δω1 = δω2∗ :
1 1p
δω1,2 = [∆1 + ∆2 + (v1 + v2 )δk] ± β1 β2 + [∆1 − ∆2 + (v1 − v2 )δk]2 . (6.57)
2 2
If δk is sufficiently large such that β1 β2 and ∆1,2 are negligible, then Eq. (6.57) leads to
1 v1 δk,
δω1,2 ≈ (v1 + v2 )δk ± |(v1 − v2 )δk| → (6.58)
2 v2 δk,
which is just the linear approximation to the unperturbed dispersion curves that correspond to δD ≈ 0.
These asymptotics cross at δk = 0; but when δk is small, the terms β1 β2 and ∆1,2 are not negligible.
To understand what happens then, consider the following. If the signs of Vω,1 and Vω,2 are opposite,
then β1 β2 = −|β|2 < 0, so the square root in Eq. (6.57) becomes imaginary at δk close enough to
69
LECTURE 6. WAVES IN COLD MAGNETIZED PLASMA 70
−(∆1 − ∆2 )/(v1 − v2 ). This signifies an instability. If the system has no free energy, though (as is
the case in cold plasma), it cannot support instabilities in principle. This guarantees that the signs
of Vω,1 and Vω,2 in such a system are the same, so β1 β2 = |β|2 > 0. Then, δω1,2 are real and
at all δk. This shows that unless β = 0 (a degenerate case), δθ induces a nonvanishing “frequency
gap” between the curves δω1 (δk) and δω2 (δk). This constitutes “level repulsion”.
70
Lecture 7
The cold-plasma model that was considered in the previous lecture misses thermal effects that can
be important. Here, we explore some of these effects semi-qualitatively within a basic fluid model. A
more accurate (kinetic) description will be presented in Part IV.
7.1 Introduction
Like in Lecture 2, let us start with the momentum equation
∂ ṽ s es 1 ∇Ps
+ (ṽ s · ∇) ṽ s = Ẽ + ṽ s × (B 0 + B̃) − + C s. (7.1)
∂t ms c ms ns
Because ∇Ps /ns is now retained, this equation must be complemented with equations for the density
ns and the pressure Ps . The density can be obtained from the continuity equation,
∂t ns + ∇ · (ns v s ) = 0, (7.2)
but handling the pressure is more complicated. A rigorous way to calculate Ps is to use the kinetic
approach (Part IV); but then the theory becomes more complicated and less transparent. Rigorous
fluid theories can be constructed asymptotically for some regimes (e.g., when plasma is strongly colli-
sional or not-so-warm), but they have limited applicability. Here, we adopt an alternative approach,
which is less rigorous but adequate qualitatively and has wider applicability.
First, consider a plasma that is in a global equilibrium. Then, as commonly done in thermody-
namics, one can assume the following simple model:
γs
ns
Ps = P0s . (7.3)
n0s
Here, the constants P0s and n0s are the pressure and density of some fixed reference state and the
constant γs is called a polytropic index. This index depends on processes of interest and can be
derived from statistical physics. For example, isotropic processes correspond to γs = 1, as seen from
the equation of state for the (ideal) gas of species s, Ps = ns Ts , where Ts is the sth-species temperature.
Adiabatic dynamics corresponds to γs = (Ds + 2)/Ds [36], where Ds is the number of relevant degrees
of freedom. Which degrees of freedom are relevant can be guessed based on qualitative arguments or
by comparing with kinetic theory.
A natural generalization of Eq. (7.3) is the model in which plasma may evolve such that each fluid
element still remains in its own local equilibrium and thus conserves its own value of Ps /nγs s . This
corresponds to conservation of Ps /nγs s in the frame moving with the fluid velocity v s ; i.e.,
d
Ps ∂ Ps
0=
d t nγss ≡ ∂t + vs · ∇ nγss . (7.4)
71
LECTURE 7. WAVES IN WARM FLUID PLASMA 72
Below, we use Eq. (7.4) along with Eqs. (7.1) and (7.2) to study linear plasma waves.1
∂ ṽ s es ∇P̃s
= Ẽ − , (7.6a)
∂t ms ms ns0
∂ P̃s ∂ ñs
= γs ms vT2 s , (7.6b)
∂t ∂t
∂ ñs
= −n0s ∇ · ṽ s , (7.6c)
∂t
.
where vT s = T0s /ms is the unperturbed thermal speed of species s.
By differentiating Eq. (7.6a) with respect to t, one obtains
∂ 2 ṽ s es ∂ Ẽ
− γs vT2 s ∇ (∇ · ṽ s ) = . (7.9)
∂t2 ms ∂t
ωp2
s = −i ω
−ω 2 1 + γs vT2 s kk j̃ (i)
Ẽ. (7.11)
4π
1 Keep in mind that the results presented below have limited applicability. As to be discussed in Part IV, some of
the waves in warm plasma are in fact heavily damped due to kinetic effects, particularly at large k.
2 Considering the general case requires cumbersome calculations, because warm plasma exhibits spatial dispersion.
72
LECTURE 7. WAVES IN WARM FLUID PLASMA 73
Let us consider projections of this equations on the axes perpendicular and parallel to k by applying
the corresponding projection matrices:
. kk . kk
Π⊥ = 1 − 2 , Πk = 2 . (7.12)
k k
(i) . (i) (i) . (i)
Since Π⊥ Πk = 0, one obtains that j̃ s⊥ = Π⊥ j̃ s and j̃ sk = Πk j̃ s satisfy
(i)
j̃ s⊥ =
i ω ωp2 Ẽ (i) iω ωp2
⊥, j̃ sk = Ẽ k . (7.13)
4π ω 2 4π ω 2 − γs k 2 vT2 s
Assuming that k is directed along the x axis, the corresponding susceptibility tensor is as follows:
ω2
− ω2 −γspk2 v2 0 0
Ts
ω2
χs (ω, k) =
0 − ωp2 0
(7.14)
ω2
0 0 − ωp2
and the dielectric tensor is
X k (ω, k) 0 0
(ω, k) = 1 + χs (ω, k) = 0 ⊥ (ω) 0 . (7.15)
s 0 0 ⊥ (ω)
The corresponding dispersion function is
k (ω, k) 0 0
D E (ω, k) = 0 ⊥ (ω) − N 2 0 , (7.16)
2
0 0 ⊥ (ω) − N
and the dispersion relation is
2
k (ω, k) ⊥ (ω) − N 2 = 0.
(7.17)
Just like in the case of cold nonmagnetized plasma (Lecture 2), there are two types of nonzero-
frequency waves in this case. One type corresponds to (Figs. 7.1 and 7.2)
N 2 = ⊥ (ω) = 0, i.e., ω 2 = ωp2 + k 2 c2 . (7.18)
These are the same transverse waves as in cold plasma (Sec. 2.2.4). Due to their transverse polar-
ization, the corresponding oscillation quiver velocities satisfy ∇ · ṽ s = 0; then, ∂t ñs = 0, and thus
∂t P̃s = 0, which is why these waves are independent of the plasma temperature.
The other type of waves predicted by Eq. (7.17) are longitudinal waves. Their dispersion relation is
X
0 = k (ω, k) = 1 + χk (ω, k) s , (7.19)
s
2
. ωps
χk (ω, k) s = − . (7.20)
ω 2 − γs k 2 vT2 s
In case of low-frequency oscillations, we expect oscillations to be isothermal; then γs = 1, so
2
ωps 1
χk,s (ω, k) ≈ ≡ 2 2 , (7.21)
k 2 vT2 s k λDs
.
where λDs = vT s /ωps is the Debye length. In contrast, high-frequency oscillations are expected to be
adiabatic and effectively one-dimensional (Ds = 1). Hence, γs = (Ds + 2)/Ds = 3, so we expect
2
ωps
χk,s (ω, k) = − . (7.22)
ω 2 − 3k 2 vT2 s
73
LECTURE 7. WAVES IN WARM FLUID PLASMA 74
N2
c2
2
3 υTi
c2
C2S
c2
3 υ2Te
ω
0 ω pi ω pe
Figure 7.1: The dispersion curves N 2 (k) for the three types of waves that can propagate in warm non-
magnetized plasma within the fluid approximation: blue – transverse electromagnetic waves, green –
electron plasma waves, red – the branch known as the ion acoustic wave (IAW) at ω ωpi and as
the ion plasma wave (IPW) at ω & ωpi .
ω ω
kc kCS
3 kυTe ω pi
ω pe
kCS
3 kυTi
ω pi
k k
0 λ-1
De 0 λ-1
Di
Figure 7.2: Left – the same as Fig. 7.1 but in coordinates ω(k). Right – a close-up of the left figure
focusing on the IAW/IPW branch and extending to larger k.
74
LECTURE 7. WAVES IN WARM FLUID PLASMA 75
ω 2 = ωpe
2
+ 3k 2 vT2 e . (7.24)
This is the dispersion relation of electron Langmuir waves, or electron plasma waves (EPW), in warm
plasma (cf. Sec. 2.2.3). Keep in mind that, by adopting γe = 3, we have restricted our model to the
high-frequency limit, ω kvT e . Then, we can as well use
q
2 + 3k 2 v 2 ≈ ω
3k 2 vT2 e
ω= ωpe Te pe + (7.25)
2ωpe
(for ω > 0). Although small, the thermal corrections are important in that they make the group
velocity nonzero:
∂ω 3kvT2 e
vg = = . (7.26)
∂k ω
Notice also that the phase and group velocities are connected by the following relation:
v g · v p = 3vT2 e . (7.27)
Another notable regime is when electrons are hot while ions are coldish. For simplicity, let us
consider a plasma with only one type of ions. Then, the corresponding dispersion relation is
2
ωpi 1
0=1− + 2 2 . (7.30)
ω2 − 3k 2 vT2 i k λDe
ω 2 ≈ CS2 k 2 , (7.32)
75
LECTURE 7. WAVES IN WARM FLUID PLASMA 76
Figure 7.3: Schematic of the electron-pressure oscillations that accompany oscillations of the cold-ion
density. Green are the ions; red are their Debye spheres, which contain electron pressure.
. 2 2 Zi T0e + 3T0i
CS2 = ωpi λDe + 3vT2 i = . (7.33)
mi
.
(Here, Zi = |ei /ee | is the ion charge state, and we have used ni0 ei + ne0 ee = 0.) These waves are
known as ion acoustic waves (IAW), or ion sound waves. Notably, they exist even in the limit of zero
Ti0 . This is understood by the fact that each ion carries a Debye cloud of electrons. Even when there
is no pressure associated with ions per se, there is an electron pressure associated with each Debye
cloud. Oscillations of the ion density cause oscillations of this electron pressure. The latter creates a
restoring force density −∇Pe , which leads to sound-like oscillations (Fig. 7.3).
At k λ−1De , Eq. (7.31) leads to
ω 2 ≈ ωpi
2
+ 3k 2 vT2 i . (7.34)
These are called ion plasma waves (IPW). The electron properties do not enter Eq. (7.34) explicitly
because sufficiently hot electrons do not contribute to k and are important for IPW only as a ho-
mogeneous neutralizing background. Also notably, the regime λ−1 −1
De k λDi corresponds to an
2 2
approximately constant frequency, ωp ≈ ωpi .
Exercise 7.1: Explore the figures ω(k) in Ref. [35] and explain, qualitatively, how they relate
to the corresponding figures for cold plasma from Lecture 6. Plot the corresponding N (ω).
76
LECTURE 7. WAVES IN WARM FLUID PLASMA 77
Figure 7.4: From Ref. [37]. For ω(k) at even larger k, see Refs. [3, 35].
77
Problems for Part III
Show that θ0 (ω) is the propagation time. Then, explain how the above result can be used to infer
n(z) from experimental measurements. What do you think are the limitations of this diagnostic?
(c) Faraday rotation. — Show that, to the lowest nonvanishing order in ω −1 , a linearly-polarized
wave propagating in plasma along a dc magnetic field experiences rotation of its polarization
angle ϑ at the rate
3 Reflection points introduce additional order-one phase shifts [cf. Eq. (2.38)], but those can be ignored here.
78
PROBLEMS FOR PART III 79
Hint: Reflection occurs when the vertical group velocity becomes zero. How does the
horizontal wave number and the frequency evolve along the rays?
Now, consider the influence of the Earth’s magnetic field B 0 , assuming it is homogeneous and parallel
to the ground. Suppose k = (k⊥ , 0, kk ), where x is the vertical axis and the z axis is along B 0 .
(b) In magnetized electron plasma, waves can experience reflection at three different locations cor-
.
responding to three different values of X = ωp2 /ω 2 . Find these values X1,2,3 from Eq. (6.30)
.
as functions of Y = |Ωe |/ω and of Nk , assuming Nk 6= 0. Assuming also that 0 < Y < 1
and Nk2 6= 1, show that there is exactly one value of Nk2 , denoted N̄k2 (Y ), at which two of the
reflection points coincide. Calculate N̄k2 (Y ) and the corresponding value of X.
4 2
(c) Show that N⊥ satisfies aN⊥ + bN⊥ + c = 0, where a, b, and c depend on X. (The dependence
on Y and Nk is also assumed.) Find a(X) and outline how to find b(X) and c(X). Without
2
solving this equation, sketch N⊥ (X) at fixed Nk2 for Nk2 ≷ N̄k2 and Nk2 = N̄k2 . You may consider
it known (or show it yourself) that
2 h
2 XY i
b − 4ac = Y 2 (Nk2 − 1)2 − 4(X − 1)Nk2 . (7.38)
1−Y2
2
Hint: Consider Nk = 0 first, for which case N⊥ (X) should be easy to find. (What are
the two modes in this regime?) Then, consider how the plot is modified for Nk 6= 0
2 2
by analyzing N⊥ (0), cutoffs, resonance(s), and the number of real roots for N⊥ .
(d) Now consider N⊥ in regions where it is real. Sketch N⊥ (X) corresponding to your sketches of
2
N⊥ (X) in part (c). (Remember to plot both N⊥ > 0 and N⊥ < 0.) Using these results, explain
the dependence of the field pattern on the launch angle α in Fig. 7.5. (Ignore the specific
numbers and focus on qualitative physics.)
79
PROBLEMS FOR PART III 80
Figure 7.5: The absolute value of the wave electric field |Ẽ(x)| (horizontal axis) versus the altitude x
(vertical axis, in km) for a standing wave launched from the ground (x = 0) at different angles α
(numbers on top) between k and the vertical. The figure is adapted from Ref. [39].
80
PROBLEMS FOR PART III 81
∂ P
+v·∇ = 0, (7.39d)
∂t ργm
where ρm is the mass density, P is pressure, and γ is a constant polytropic index. As usual, assume
that the background plasma is homogeneous and has no average velocity. Also assume homogeneous
stationary magnetic field B 0 . For simplicity, you may also adopt the coordinate system such that
B 0 = B0 ēz and k = k⊥ ēx + kk ēz . (Here, ēa is a unit vector along ath axis.)
(a) Linearize Eqs. (7.39) and derive a PDE for the fluid velocity; then use it to derive the dispersion
relation by adopting ∂t = −i ω and ∇ = i k. Show that this dispersion relation can be represented
in the form
(b) Plot in polar coordinates the phase speed as a function of the angle between k and B 0 for all
branches. Which of these waves can propagate along B 0 ? Which of these waves can propagate
perpendicularly to B 0 ?
(c) Approximate ω 2 for VS VA . What is the physical mechanism of each branch in this limit?
(d) Calculate the energy density of an Alfvén wave in a cold plasma with negligible kk in the limit
γA 1. (The answer is the same for both types of Alfvén waves.)
81
Part IV
In this part of the course, we extend our previous models of plasma waves by including
kinetic effects.
82
Lecture 8
Fluid theory used in the previous lectures is only a rough reduction of the complete (Klimontovich)
description that accounts for each individual particle in the plasma. The next best approach is the
kinetic approach, in which plasma is considered as a fluid in phase space. This lecture is intended as
an introduction into kinetic theory and its applications to modeling plasma waves.
8.1 Introduction
8.1.1 Distribution function
Suppose that the motion of a single particle is fully described by some set of phase-space variables Γ,
for example, (x, v) or (x, p). Then, one can define the particle phase-space density F (Γ) in space Γ,
independently for each given species (Box 8.1). Let us consider two different sets of such variables,
assuming that they are connected by some invertible function Q,
Q : Γ1 7→ Γ2 . (8.1)
First, suppose the phase-space density corresponding to a single particle with some coordinate Γ20 .
Such phase-space density is given by δ(Γ2 − Γ20 ). In terms of the particle coordinate Γ10 in the Γ1
space, the same function can be expressed as δ[Γ2 − Q(Γ10 )]. In the case of multiple particles, the
corresponding phase-space density in Γ2 must be averaged over all Γ10 , or in other words, integrated
over their phase-space density F 1 . This gives the following general rule for mapping the phase-space
density F 1 in Γ1 to the phase-space density F 2 in Γ2 :
F 2 (Γ2 ) = d Γ1 δ[Γ2 − Q(Γ1 )]F 1 (Γ1 ) = F 1 [Q−1 (Γ2 )] ∂Γ1
∂Γ2
, (8.2)
where the latter ratio denotes the Jacobian of the corresponding variable transformation and Q−1 is
the function inverse to Q.
Among all possible variable transformations, there are so-called canonical transformations, which
are special. They correspond to unit Jacobians, so the canonical phase-space densities (which we
denote with F as opposed to the general phase-space densities F ) are transformed simply as
In this sense, the canonical phase-space density F is an invariant with respect to canonical transfor-
mations.
83
LECTURE 8. INTRODUCTION TO KINETIC THEORY OF PLASMA WAVES 84
Box 8.1: Kinetic modeling of quantum plasmas and general broadband waves
Here, we discuss only classical kinetic theory. For quantum particles, phase-space coordinates are
operators and there is no such thing as the phase-space density, so the theory has to be formulated
differently. For a system described by a state function |ψi, one starts by introducing the density
.
operator %̂ = |ψi hψ| and its Weyl symbol %, which is known as the Wigner function (or Wigner
tensor, if hx|ψi is a vector). The Wigner function satisfies the Moyal equation ∂t % = {{H, %}},
where H is the Weyl symbol of the particle Hamiltonian and {{... , ... }} are the so-called Moyal
brackets [40]. This equation is a generalization of the kinetic equation that is derived below.
(The classical distribution function is the coarse-grained limit of the Wigner function % up to a
constant factor.) The same formalism is applicable to classical waves [41, 42].
0=
d F [t, x(t, x , p ), p(t, x , p )] = ∂Fs + d x · ∂Fs + d p · ∂Fs , (8.7)
dt s 0 0 0 0
∂t d t ∂x d t ∂p
which is known as Liouville’s theorem (Box 8.2).
In nonrelativistic plasma physics, it is customary to work with the particle phase-space density
f (t, x, v) in the (x, v) space instead of the canonical phase-space density F(t, x, p). The relation
between f and F is obtained using Eq. (8.2):
∂(x, p)
fs (t, x, v) = Fs [t, x, p(t, x, v)] . (8.8)
∂(x, v)
For nonrelativistic plasmas, which we consider below, one can take
es
p(t, x, v) = ms v − A(t, x), (8.9)
c
where A is the electromagnetic vector potential. The corresponding Jacobian is
∂(x, p)
= m3s , (8.10)
∂(x, v)
so fs and Fs differ only by a constant factor. Therefore, they satisfy the same equation,
∂fs F s ∂fs
+ v · ∇fs + · = 0, (8.11)
∂t ms ∂v
where F s = ms v̇ is the force.
84
LECTURE 8. INTRODUCTION TO KINETIC THEORY OF PLASMA WAVES 85
Liouville’s theorem can also be interpreted as follows. Consider a small fluid element in phase
space. The shape of this element may become distorted with time but its volume d Γ is conserved,
because the transformation (8.4) is canonical. The number of particles inside this volume is
conserved too, by definition, so Fs d Γ = const. Then, Fs = const along the trajectory followed
by each given phase-space element.
85
LECTURE 8. INTRODUCTION TO KINETIC THEORY OF PLASMA WAVES 86
.
where terms quadratic in the wave field have been dropped and Ωs = es B 0 /(ms c), as usual. (Unlike in
the fluid description that we studied earlier, the velocity is an independent variable here; thus, v · ∇f˜s
is a linear term and thus must be retained.) The first-order magnetic field that enters Eq. (8.17) can
be expressed through Ẽ using Faraday’s law:
For simplicity, we will limit our consideration to homogeneous plasmas and to waves of the form
∝ e i k·x . Then, k̂ can be replaced with k. However, the time dependence has to be handled in a more
subtle manner, as will be discussed in Lecture 9.
∂ f˜s
+ i kvx f˜s = −
es n0s ∂f0s
Ẽ , (8.23)
∂t ms ∂vx
where f˜s and Ẽ are now independent of x but depend on k, and the x axis is chosen parallel to k.
This equation is similar, for example, to Eqs. (2.5) and (6.4) and can be solved in the same way:
s (v)e
−i kvx t
f˜s = f˜(f) + f˜(i)
s (t, v), (8.24)
t
f˜(i)
s =−
es n0s ∂f0s (v)
ms ∂vx
d t0 e −i kv 0
x (t−t )
Ẽ(t0 ), (8.25)
0
(f)
where f˜s is determined by initial conditions but not by the field.
86
LECTURE 8. INTRODUCTION TO KINETIC THEORY OF PLASMA WAVES 87
f˜(i)
s (t, v) = −
ms ∂vx
e
es n0s ∂f0s (v) −i kvx t
. (8.26)
(This is understood as the Green’s function of the linearized Vlasov equation.) The corresponding
∞
induced current es −∞ d v vx f˜s (t, v) is, by definition, the conductivity in the (t, k) representation
(i)
[which is the Fourier image of the function Σ̄s (t, x) that we used earlier]:
∞
e2s n0s
Σ̄s,k (t) = −
ms
d v vx ∂f∂v e
0s (v) −i kv xt
. (8.27)
−∞ x
87
Lecture 9
Here, we discuss the concepts of eigenmodes in kinetic theory and dispersion relation in kinetic theory,
which are more subtle than those in fluid theory.
Let us complement this equation with Ampere’s law. For electrostatic oscillations, it can be written
as ∂t Ẽ = −4π j̃x , or equivalently,
∂t Ẽ = −4πee d vx vx f˜e . (9.3)
By using a variable transformation f˜e → i αf˜e with an appropriate real constant α, one can bring this
set of equations to the form
i ∂t f˜e = kvx f˜e + vx Ẽ, i ∂t Ẽ = d vx vx f˜e . (9.4)
Let us discretize the velocity space into N chunks of the size ∆v centered√around vx = va ,
.
a = 1, 2, . . . , N . Let us also replace f˜e → (∆v)−1/2 f˜, and introduce ga = va ∆v. This brings
Eq. (9.4) to the Schrödinger form
i ∂t ψ = Hψ, (9.5)
88
LECTURE 9. EIGENMODES IN KINETIC THEORY 89
Hermitian Hamiltonians are often represented as graphs. Graph’s nodes correspond to the com-
ponents of the state vector ψ and edges denote nonzero elements of the Hamiltonian matrix. In
particular, each diagonal element connects the corresponding node with itself, so is it represented
as a loop. The Hamiltonian (9.6) corresponds to a graph that is a “star with loops”: the node
that represents Ẽ is connected with each node representing f˜a , but no f˜a and f˜b are connected
directly with each other unless a = b. Such topology of Hamiltonian graphs is a signature feature
of collisionless-plasma models and mean-field theories in general.
where ψ = (f˜1 , f˜2 , . . . , f˜N , Ẽ)| is a column vector and (Box 9.1)
kv1 0 0 ... 0 g1
0 kv2 0 . . . 0 g2
0 0 kv 3 . . . 0 g3
H= ... ... ... ... ...
. (9.6)
...
0 0 0 . . . kvN gN
g1 g2 g3 . . . gN 0
Let us search for eigenmodes of this system in the form ψ ∝ e −i ωt . The corresponding frequencies ω
satisfy the following equation (Exercise 9.1):
N
X ga2
0 = D(ω) = ω − . (9.7)
a=1
ω − kva
Exercise 9.1: Derive Eq. (9.7) by proving the following equality for general ϑa and ga :
ϑ1 0 0 ... 0 g1
0 ϑ2 0 ... 0 g2 ! N
N
|ga |2 Y
0 0 ϑ3 ... 0 g3 X
det = ϑ− ϑb .
... ... ... ... ... ... ϑa
a=1 b=1
0 0 0 . . . ϑN gN
g1∗ g2∗ g3∗ ∗
. . . gN ϑ
As an algebraic equation of order N + 1, Eq. (9.7) has N + 1 solutions for ω. All these solutions are
real because H is Hermitian (Exercise 9.2). The fact that Eq. (9.7) always has N + 1 real solutions
can also be seen graphically as shown in Fig. 9.1. The corresponding eigenvectors satisfy
ga
f˜a = Ẽ. (9.8)
ω − kva
Exercise 9.2: Show that if f¯(E) is nonmonotonic, then H is not Hermitian and that the shape
of D(ω) changes qualitatively such that complex roots in general become possible.
In the limit N → ∞, these modes are known as Case–van Kampen modes [44, 45].1 The mode
spectrum becomes continuous in this case, so any ω is an eigenfrequency. Strictly speaking, the concept
1 Here, we consider only a special case of Case–van Kampen modes that corresponds to monotonic f¯(E). One can
also generalize these modes by adding ion motion and background magnetic field.
89
LECTURE 9. EIGENMODES IN KINETIC THEORY 90
Figure 9.1: A graphical solution of Eq. (9.7) for N = 4. The solutions (black points) correspond to
the crossings of D of the horizontal axis. The vertical dashed lines correspond to ω = kva , the oblique
line represents the asymptotic D(ω) = ω.
of a dispersion relation becomes irrelevant then. In fact, for any bounded absolutely integrable function
u, one can find initial conditions such that Ẽ(t) = u(t) [46]. The only subtlety is that such initial
conditions are typically not analytic [because the eigenvectors (9.8) are not analytic] and thus are not
realized naturally. To study what happens for “natural” initial conditions, one should consider the
general initial-value problem. This will be discussed in Lecture 9.
where the system is assumed homogeneous. It can be converted into a simpler, algebraic, equation
by applying the Fourier transform in x and the Laplace transform in time. Originally, the Laplace
transform of a given function F is defined as
∞
LF :
.
F (t) 7→ F̄ (s) = d t e −st F (t), (9.11)
0
90
LECTURE 9. EIGENMODES IN KINETIC THEORY 91
The convolution theorem for the Laplace transform is valid only if the corresponding functions
are sufficiently well behaved. The dispersion operator D̂ E in the form (9.9) integrates the field,
so it is well behaved. In contrast, differential operators have singular kernels and require special
treatment. In particular,
(and so on for higher-order derivatives), as can be seen via taking the integral in Eq. (9.13) by
parts. This means that if D̂ is a polynomial of ∂t , then D̂ Ẽ = S becomes D̄(ω, k)Ē(ω, k) =
S̄ + ∆S̄ , where D̄(ω, k) and ∆S̄ are polynomials of ω and ∆S̄ is determined by F (t0 ), F 0 (t0 ), ...
In cold-plasma problems, it is convenient to work with the field equation in the form
4π i ω̂ (f) ω̂ 2
D̂ Ẽ = − j̃ , D̂ = D̂ E ,
c2 c2
(f)
because f˜s are time-independent in the absence of thermal motion and thus ω̂ j̃ (f) = 0. In
this case, the spectral representation of the field equation is simply D̄(ω, k)Ē(ω, k) = ∆S̄ . But
because this equation has the same form as Eq. (9.15) up to the definition of the dispersion
operator and the source term, the discussion of the initial-value problem in the main text applies
to such problems just as well.
where Re s must be large enough for the integral to converge. The inverse transform is
a+i b
L−1 F̄ : F̄ (s) 7→ F (t) =
1
lim
2π i b→∞
d s e st F̄ (s), (9.12)
a−i b
where a is a real number that can be chosen arbitrarily as long as it is large enough such that all
singularities of F̄ remain to the left from the contour B. Here, we assume an equivalent but a slightly
modified definition:
∞
F̄ (ω) = d t e i ωt F (t), (9.13)
t0
+∞+i a
F (t) =
1
2π
d ω e −i ωt F̄ (ω), (9.14)
−∞+i a
where the former integral is now taken from t0 instead of zero and a variable transformation s = −i ω
has been applied. The latter makes the Laplace transform look similar to the Fourier transform, except
for the following: (i) the integral in Eq. (9.13) is taken over half of the time axis; (ii) ω is generally
complex and must have a large enough imaginary part for the integral in Eq. (9.13) to converge; and
(iii) the integral in Eq. (9.14) is taken not necessarily along the real axis but parallel to it, specifically,
at a distance a such that all singularities of the integrand remain below the contour.
The Laplace transform in time and the Fourier transform in space convert the convolution integral
in Eq. (9.10) into a product (Box 9.2), so one obtains
91
LECTURE 9. EIGENMODES IN KINETIC THEORY 92
Then, the contour sticks only to singularities of D −1E . By definition of an inverse matrix, it follows
that such singularities are possible, in particular, when
If det D E is analytic near such points, it can be Taylor-expanded. Then, the corresponding singular-
ities of D −1
E are poles. We denote them as ωq (k).
Note now that, as we are shifting the contour further toward Im ω → −∞, the exponent under the
integral approaches zero (at t > 0), because
Assumption 3: Let us assume that D −1 E S does not grow too rapidly (if at all) at
Im ω → −∞, so the smallness of the aforementioned exponent is enough to ensure that
the integral over the horizontal part of the contour vanishes.
Under these assumptions, the whole integral can be expressed as the sum over the contributions
of the aforementioned poles only, i.e.,2
Rq (k)e −i ωq (k)t ,
X
Ẽ k (t) = (9.19)
q
where the coefficients Rq are determined by the initial conditions, or more specifically, proportional
E (ω, k)S (ω, k) at ω = ωq (k). In principle, one can choose initial
to the corresponding residues of D −1
conditions such that all but one of these coefficients be zero; then, the resulting field is
Because this solution has a well-defined complex frequency, ωq (k), it can be considered as an eigen-
mode. Also, the mode polarization Rq can in general be found as follows. By applying the Laplace
transform to Eq. (9.20), we obtain
Ẽ(ω, k) =
i Rq (k) . (9.21)
ω − ωq (k)
in Eq. (9.19), with m = 1, 2, . . . (n − 1). This case will not be considered because it is not typical. In particular, the
presence of high-order zeros of det D E does not necessarily imply the presence of high-order poles in D −1 E .
92
LECTURE 9. EIGENMODES IN KINETIC THEORY 93
93
Lecture 10
Dispersion properties of
nonmagnetized plasma
Here, we apply the results of the previous lectures to study the dispersion properties of nonmagnetized
plasma.
i
˜(i) es n0s ∂f0s v·k vb kc
fs = − δcb 1 − + Ẽb . (10.1)
ms (ω − k · v) ∂vc ω ω
94
LECTURE 10. DISPERSION PROPERTIES OF NONMAGNETIZED PLASMA 95
Lemma 1
where C is some contour in the complex-v space and h is an arbitrary function. Assuming
that h is absolutely integrable on C, it is easy to see that IC (u) is well-defined for all
u∈/ C, and so are all its derivatives. Therefore, IC (u) is analytic for all u ∈
/ C.
Lemma 2
Suppose that u is initially in the region A1 that is above a given contour C. Consider
moving u to the region A2 that is on the other side of the contour C. When u crosses
C, the function IC (u) may become singular. But let us consider the integral IL (u) taken
over contour L that is the same as C except it is bended to remain below the pole at all v.
The is called a Landau contour. Then, according to Lemma 1, the function IL (u) remains
analytic. Because one also has IL (u) = IC (u) in the whole region A1 , the function IL (u)
represents the analytic continuation of IC (u) to the region A2 (and remember that the
analytic continuation is unique). By splitting the integral into the principal-value part and
the pole contribution, one finds that the expression
0, u ∈ A1 ,
d v vh(v)
−u
= d v
h(v)
v−u
+ i πh(u) × 1, u ∈ C, (10.6)
L C
2, u ∈ A2
represents the analytic continuation of IC (u). [Both terms in Eq. (10.6)are discontinuous
but their sum is continuous.] Note that unless u ∈ C, one can replace with , because
then the principal value of the integral coincides with the integral itself. Also note that
if the initial region A1 corresponds to the region below C, then i π in the above formula
must be replaced with −i π.
This result may seem surprising in the following sense. The same formula for Im ω > 0 is read-
ily obtained from the linearized Vlasov equation if one simply adopts ∂t = −i ω. But for strictly
monochromatic oscillations of the distribution function, ∂t f˜s = −i ω f˜s is a precise equality irrespec-
tive of the sign of Im ω, so why should one expect a different result? The explanation is as follows.
Remember that f˜ consists of two terms: the field-induced term f˜s ∝ exp(−i ωt) and the term f˜s
(i) (f)
that is determined by free oscillations, which has the time dependence of the form f˜s ∝ exp(−i kvt).
(f)
(i) (f)
At Im ω > 0, f˜s eventually becomes much larger than f˜s , so the initial conditions do not matter.
(More precisely, the moment t0 at which the initial conditions are prescribed can be shifted to −∞,
(i) (i)
so f˜s = f˜s at any finite t.) But at Im ω < 0, f˜s is essential, and
∂t f˜s ≈ −i ω f˜(i)
s − i kv f s 6= −i ω fs .
˜(f) ˜ (10.8)
Although such f˜s is not monochromatic, integrals of this function (such as the current density) can
be monochromatic, which is how plasma supports the quasimodes that were introduced in Lecture 9.
95
LECTURE 10. DISPERSION PROPERTIES OF NONMAGNETIZED PLASMA 96
. 1 X 2 . X 2
f0 (v) = 2 ω f0s (v), ωp2 = ωps , (10.9)
ωp s ps s
P
and using ab (ω, k) = δab + s [χab (ω, k)]s with Eq. (10.7), one obtains
ωp2
k·v
ab (ω, k) = δab +
ω
d v ω − k · v δcb 1 − ω + ω ∂v .
va kc vb ∂f0
(10.10)
L c
Because
∞
δab ∞
k · v ∂f0s
d v ω − k · v δcb 1 − ω ∂v = ω1
va
d v va ∂f0s
∂vb
=−
ω −∞
d v f0s = −
δab
ω
, (10.11)
L c −∞
one can also present this in the following alternative form, which will be useful below:
ωp2 ωp2
ab (ω, k) = 1 − 2 δab + 2
ω ω L
d v
kc va vb ∂f0
ω − k · v ∂vc
. (10.12)
∂ ∂ q 2 va ¯0
f0 (v) = f¯00 (v) vx + vy2 + vz2 = f (v), (10.13)
∂va ∂va v 0
and it is easy to see that that acquires the following diagonal form:
k 0 0
= 0 ⊥ 0 , (10.14)
0 0 ⊥
assuming that the x axis is chosen along k. Correspondingly, the dispersion matrix is as follows:
k (ω, k) 0 0
D E (ω, k) = 0 ⊥ (ω, k) − N 2 0 , (10.15)
2
0 0 ⊥ (ω, k) − N
.
where N = kc/ω. This means that like in fluid plasma, waves can be of two types: transverse
electromagnetic waves and longitudinal electrostatic waves.
96
LECTURE 10. DISPERSION PROPERTIES OF NONMAGNETIZED PLASMA 97
This is similar to the electromagnetic-wave dispersion that we studied earlier except ⊥ (ω, k) is some-
.
what different from its cold limit cold (ω) = 1 − ωp2 /ω 2 . From Eq. (10.12), one has
ωp2 ωp2
ab (ω, k) = 1 − 2 δab + 2
ω ω L
d v ωkv−akv
vb ∂f0s
, (10.17)
x ∂vx
(Here, we have assumed that ω/k & c by analogy with waves in cold plasma, and we have also assumed
vT c, because our theory is nonrelativistic to begin with.) Then,
This indicates that the electromagnetic waves in isotropic nonmagnetized plasma are not significantly
affected by (nonrelativistic) temperature; i.e., their dispersion relation is approximately the same as
in cold plasma,
ω 2 ≈ ωp2 + k 2 c2 . (10.20)
and
2
ωps
χs,k (ω, k) =
ω
d v ω − kv δxx 1 − ω + ω ∂v .
vx kvx kvx ∂f0s
(10.22)
L x x
Let us integrate over vy and vz and change the notation as in Eq. (8.28). Then,
2
ωps
χs,k (ω, k) =
ω
d v ω −v kv ∂f∂v0s , (10.23)
L
Then,
0
f 0 (v)
d v vf0s (vx )
ω − kv
= dv 1 1
− − 2
ω
k k v − ω/k
0
f0s
ω
(v) = − 2
k L
d v 0s
v − ω/k
, (10.25)
L x L
∞
where we used that −∞ d v f0s 0
(v) = f0s (+∞) − f0s (−∞) = 0. This leads to the following expression
for the susceptibility (Exercise 10.1):
2 0
ωps
χs,k (ω, k) = − 2
k
d v vf−0s (v)
ω/k
, (10.26)
L
97
LECTURE 10. DISPERSION PROPERTIES OF NONMAGNETIZED PLASMA 98
∞
Exercise 10.1: Show that Eq. (10.26) can as well be derived using σs (ω, k) = 0
d t e i ωt Σs,k (t),
with Σs,k taken from Eq. (8.29).
or more explicitly,
2
ωps ∞ 0 2
ωps ω 0, Im ω > 0,
χk,s (ω, k) = − 2
k
d v v − ω/k − i π k|k| f0s0 k × 1, Im ω = 0,
f0s (v)
(10.27)
−∞ 2, Im ω < 0.
Because v and the phase velocity ω/k can be comparable, the influence of kinetic effects on the
susceptibility can be strong, so let us discuss longitudinal waves in more detail. For simplicity, let us
focus on waves with small ωi and k (ω, k) that is smooth enough. Then, Eq. (10.21) can be simplified
as follows:
Because k on the right-hand side is evaluated at ωr , it can be calculated using Eq. (10.27) with
Im ω = 0:
The real and imaginary part of this complex equation give equations for ωr and ωi ; specifically,2
r (ωr , k) = 0, (10.31a)
i (ωr , k)
ωi = − . (10.31b)
∂ω r (ωr , k)
In Lecture 11, we will apply these results to study longitudinal waves in Maxwellian plasma.
(ω) = 0, (10.32)
where the index k and the argument k are omitted for brevity and
ωp2 0
. X ωps
2
(ω) = 1 − 2
k
d v v f−0 (v)
ω/k
, f0 (v) =
ωp2
f0s (v). (10.33)
L s
2 Remember that this model holds only for smooth enough (ω, k). Distributions with narrows beams can be
k
unstable even when i = 0; for example, see Problem PI.3.
98
LECTURE 10. DISPERSION PROPERTIES OF NONMAGNETIZED PLASMA 99
At Im ω > 0, the Landau contour L in Eq. (10.33) can be replaced with the real axis:
∞
ωp2 0
(ω) = 1 − 2
k
d v v f−0 (v)
ω/k
, (10.34)
−∞
where the argument k is omitted for brevity. If f00 is absolutely integrable, the right-hand side is
analytic, which is proven like Lemma 1 in Lecture 10. At Im ω = 0, one has
ωp2 ∞ 0 ω2
d v v f−0 (v) − iπ
ω
p
(ω) = 1 − f0 . (10.35)
k2 −∞ ω/k k|k| 0 k
This function is also analytic provided that f0 is a physical distribution, which are always smooth on
the real axis.3 Thus, (ω) is analytic at Im ω ≥ 0. This property allows one to assess plasma stability
without actually solving Eq. (10.32), specifically, as follows.
At large enough ω, one has σ ∝ ω −1 (Appendix AI.1), so 0 (ω) ∝ ω −3 . Thus one can just as well
write Eq. (10.36) as
0
N̄ =
1
2π i
d ω (ω)
(ω)
, (10.37)
C
where C is the closed contour that includes the real axis R and a semi-circle of infinite radius at
Im ω > 0. By analyticity of in the area encircled by C, the function 0 (ω)/(ω) can have singularities
within C only at ω = ωq that satisfy (ωq ) = 0. Also due to analyticity of , one has (ω) = αq (ω−ωq )nq
in the vicinity of these points, where αq is some constant, and nq is some positive integer. Then, the
integral over C can be expressed as a sum of residues at ωq :
0
d ω (ω)
X 1 (ω)
N̄ =
q
2π i Cq
nq −1
− ωq )
d ω nqααq (ω
X 1
=
q
2π i Cq (ω − ω )
q
n
q
q
d ω ω −1 ω
X nq
=
q
2π i Cq q
X
= nq . (10.38)
q
In the case of simple poles (nq = 1), the right-hand side equals the sum of zeros of in the upper half
of the complex-ω plane, i.e., the number of unstable modes. One can also extend this statement to
general nq assuming the convention that each nq th-order zero counts as nq zeros.
Notice now that N̄ can also be calculated differently:
N̄ =
1
d ω
0 (ω)
=
1 d = 1
[ln (ω → +∞) − ln (ω → −∞)], (10.39)
2π i R (ω) 2π i (R) 2π i
3 In contrast, away from the real axis, f (v) is not necessarily analytic (cf. Problem PIV.2), so this argument is not
0
extendable to Im ω < 0.
99
LECTURE 10. DISPERSION PROPERTIES OF NONMAGNETIZED PLASMA 100
where (R) is the projection of the real axis in the ω space to the space. By definition of the complex
logarithm, ln = ln || + i ϑ, where ϑ = arg . Also, |(ω → ±∞)| = 1, so
.
Theorem: The number of unstable modes equals the number of times the contour (R)
encircles the origin.
ωp2 ∞ 0 ω2
d v v f−0 (v) − i π 2 f00
ω
p
(ω) = 1 − . (10.41)
k2 −∞ ω/k k k
Let us assume that, when projected to the complex- plane, this contour encircles the origin. For this
to occur, there must be some ω = ω∗ such that Re (ω∗ ) < 0 and Im (ω∗ ) = 0. Since ω∗ is real by
definition of R, this requires the following two conditions be satisfied simultaneously:
ωp2 ∞ 0
1−
k2
d v v −f0ω(v)/k < 0, (10.42a)
−∞ ∗
ωp2 0 ω
∗
π 2 f0 = 0. (10.42b)
k k
Suppose that f0 (v) has only one peak, say, at some v = v∗ . Then, Eq. (10.42b) implies ω∗ /k = v∗ ,
so the inequality (10.42a) can be expressed as follows:
ωp2 ∞ 0
k2
d v vf0−(v)v > 1. (10.43)
−∞ ∗
But because the signs of f00 (v) and of (v − v∗ ) are opposite at all v, this integral is negative; hence,
the inequality (10.43) cannot be satisfied. This means that our assumption regarding the existence of
ω∗ is invalid, so (R) cannot encircle the origin. This means that by the Nyquist theorem, single-peak
distributions cannot support unstable modes.
4 In the figures below, which are taken from Ref. [1], H is the same as our .
100
LECTURE 10. DISPERSION PROPERTIES OF NONMAGNETIZED PLASMA 101
However, as long as v1 and v3 are not too far from each other, the loop is small and thus does not
encircle the origin. This means that by the Nyquist theorem, the appearance of a second peak does
not immediately lead to an instability. Having an instability requires the peaks to be sufficiently far
from each other.
101
Lecture 11
In this lecture, we apply the results of the previous lectures to study basic electrostatic waves in
isotropic Maxwellian plasma. (Because only such waves will be considered, the index k will be omitted.)
First, we present two alternative but equivalent calculations of the Maxwellian-species susceptibility.
Then, we discuss asymptotic approximations and, finally, we apply them to explicitly calculate an
approximate dispersion relation for Langmuir waves and ion acoustic waves in Maxwellian electron
plasma.
v2
1
f0s (v) = √ exp − 2 . (11.1)
2πvT s 2vT s
102
LECTURE 11. ELECTROSTATIC WAVES IN ISOTROPIC MAXWELLIAN PLASMA 103
0.5
kυTs t
1 2 3 4 5
-0.5
Using
(αe −α /2 ) = (1 − α2 )e −α /2 ,
∂ ∂ 2 2
[αI(α)] = (11.4)
∂α ∂α
one obtains
2
ωps (kvT s t)2
2
Σ̄s,k (t) = 1 − (kvT s t) exp − . (11.5)
4π 2
−1
It is seen then that the phase mixing occurs on the time scale (kvT s ) , as anticipated (Fig. 11.1).
Also, using Eq. (11.5), one can readily calculate the conductivity in the spectral representation,
2 ∞
ωps
d t 1 − (kvT s t) exp i ωt − 2 (kvT s t) .
1
h i
2 2
σs (ω, k) = (11.6)
4π 0
It is also common to express this result as follows. Let us introduce the so-called plasma dispersion
function 1 (Fig. 11.2)
∞
z2 √
Z(ζ) = i dz exp i ζz − 4 = i πe −ζ 2 − 2S(ζ),
.
(11.7)
0
4π i σs (ω, k) 2
ωps Z 0 (ζs )
χs (ω, k) = = − 2 ζs2 Z 0 (ζs ) = − 2 2 , (11.10)
ω ω 2k λDs
.
where λDs = vT s /ωps is the Debye length of species s. Because Z is an entire function of its argument,
the expression on the right-hand side of Eq. (11.10) is an entire function of ω. In particular, this means
that this expression can be used at any Im ω.
1 We assuming k > 0 for simplicity. For the general case and additional details, see Sec. 8.14 in Ref. [1]. Some
k
properties of the plasma dispersion function are also summarized in Appendix AIV.1.
103
LECTURE 11. ELECTROSTATIC WAVES IN ISOTROPIC MAXWELLIAN PLASMA 104
Figure 11.2: Re Z(ζ) (left) and Im Z(ζ) (right) as functions of (Re ζ, Im ζ).
Note that
e −z = √1 de −z
2 2
!
√
1
π
dz −2z
z − ζs π
dz z −1 ζ dz
L L s
−z 2 d
= −√
1
π L
dz e dz z − ζ 1
s
2 d
=√
1
π L
dz e d ζ z − ζ
−z 1
s s
1 d
=√
π d ζs L
dz e −z 2 1
z − ζs
. (11.13)
Then, finally,
dz ze − ζ .
2
Z̃ 0 (ζs ) . 1 −z
χs (ω, k) = − 2 2 , Z̃(ζ) = √ (11.14)
2k λDs π L
104
LECTURE 11. ELECTROSTATIC WAVES IN ISOTROPIC MAXWELLIAN PLASMA 105
This result is identical to Eq. (11.10) except the plasma dispersion function Z is replaced with Z̃.
Thus, Z̃ given by (11.14) is just an alternative representation of Z.
dz ze − ζ ,
∞ 2
−z √
π e −ζ .
1 2
Zr (ζ) = √ Zi (ζ) = (11.17)
π −∞
Then,
X Z 0 (ζr )
r
r (ωr , k) = 1 − , (11.18a)
s
2k 2 λ2D s
√ X ζr e −ζ
2
X Z 0 (ζr )
i
r
i (ωr , k) = − = π , (11.18b)
s
2k 2 λ2D s s
k 2 λ2D s
√
.
where ζr = ωr /(kvT 2). The index r in ζr is henceforth dropped for brevity.
11.2 Asymptotics
11.2.1 Warm species
Let us search for waves in plasma with “warm” species, i.e., such that ζ is large but finite. Then, due
to the exponential factor, the integral is mainly determined by the integrand at z ζ. Then,
∞
1 1 1 X zn
=− ≈− . (11.19)
z−ζ ζ(1 − z/ζ) ζ n=0 ζ n
This leads to
∞ ∞ ∞ ∞
ζ −1 zn ζ −1 z 2n
dz e −z 2
dz e −z 2
X X
Zr (ζ) = − √ = − √ , (11.20)
π −∞ n=0
ζn π −∞ n=0
ζ 2n
where we used the fact that the integrals with odd powers of z are zero. Hence,
∞
1 1
Zr (ζ) ≈ − J(0) + 2 J(1) + . . . ,
ζ ζ
. 1
J(n) = √
π −∞
d z z 2n e −z .
2
(11.21)
This integral can be expressed through the gamma function. Alternatively, notice that
105
LECTURE 11. ELECTROSTATIC WAVES IN ISOTROPIC MAXWELLIAN PLASMA 106
(n)
where denotes the nthe-order derivative and
∞ ∞
¯
J(β)
. 1
=√
π
dz e −βz 2
=√
1
dz e −z 2 1
=√ . (11.23)
−∞ πβ −∞ β
Then, one readily obtains
¯ = 1,
J(0) = J(1) J(1) = −J¯0 (1) = 1/2, (11.24)
Using Eq. (11.15), one then obtains the real part of the susceptibility in the form
1 1 3
χs,r (ωr , k) ≈ − 2 2 + 4
2k λDs ζs2 2ζs
2 2 2
ωps 2k vT s
3
=− 2 2 1+ 2
2k vT s ωr2 2ζs
2 2 2
ωps
3k vT s
=− 2 1+ (11.26)
ωr ωr2
and the imaginary part of the susceptibility in the form
√ ζs e −ζs
2
χs,i (ωr , k) ≈ π 2 2
k λDs
r 2
π ωr ωps
=
2 kvT s k vT2 s
2
e −ζs2
π ωr e −ζs
r 2
= . (11.27)
2 ωps (kλDs )3
It is also instructive to notice that Re Z(ζ) = −2ζ for real ζ is seen from the following:
d z ze − ζ du e
2
∞ −z ∞ −(ζ+u)2
1 1
√ =√
π −∞ π −∞ u
= √
e −ζ 2 ∞
du e
−u2 −2ζu
e
π −∞ u
2 The second term on the right-hand side in Eq. (11.29) is much smaller than the first term and may, in fact, be
comparable to the corrections that we neglected in the Dawson function approximation. However, this term is more
important as it is the primary cause of damping, whereas corrections to the Dawson function would only slightly affect
the real part of the dispersion relation.
106
LECTURE 11. ELECTROSTATIC WAVES IN ISOTROPIC MAXWELLIAN PLASMA 107
du e u
∞ −u2
1
≈√ (1 − 2ζu)
π −∞
du e
∞ 2 ∞
−u
=√
1
π u
2ζ
−√
π
d u e −u
2
−∞ −∞
= −2ζ. (11.30)
11.3 Waves
Now let us use the above results to derive the dispersion relations for Langmuir wave and ion acoustic
from kinetic theory.
π ωr e −ζe
2 2
3k 2 vT2 e
r
ωpe
r (ωr , k) ≈ 1 − 2 1 + , (ω
i r , k) ≈ . (11.31)
ωr ωr2 2 ωpe (kλDe )3
Equation (10.31a) leads to
2
ωpe 3k 2 vT2 e
0 = r (ωr , k) ≈ 1 − 2 1 + , (11.32)
ωr ωr2
2
so one obtains ωr2 = ωpe + O k 2 vT2 e . The second term on the right-hand side is small, so to the
lowest (zeroth) order in the temperature, one has ωr2 ≈ ωpe 2
, as expected. Then, to the next (first)
order in the temperature, one has
2
ωpe 3k 2 vT2 e
0=1− 2 1+ 2
. (11.33)
ωr ωpe
This leads to
3k 2 vT2 e
2 2 2
ωr = ωpe 1+ 2
= ωpe + 3k 2 vT2 e , (11.34)
ωpe
or in other words,
3
ωr ≈ ±ωpe 1 + k 2 λ2De . (11.35)
2
Equation (10.31b) leads to
π |ωr |3 e −ζe
r 2
ωi 1 i (ωr , k)
=− ≈− 3 3, (11.36)
|ωr | ωr ∂ω r (ωr , k) 2 2ωpe (kλDe )
where we have substituted
2 2
ωpe 2ωpe
∂r (ωr , k) ∂
≈ 1− 2 = 3 . (11.37)
∂ω ∂ω ωr ωr
2
We can also use the approximation |ωr | ≈ ωpe but ζe in e−ζe must be calculated more accurately:
2
ωr2 ωpe + 3k 2 vT2 e 1 3
ζe2 = = = 2+ , (11.38)
2k 2 vT2 e 2k 2 vT2 e 2κ 2
107
LECTURE 11. ELECTROSTATIC WAVES IN ISOTROPIC MAXWELLIAN PLASMA 108
2.0
0.100
1.5
0.010
-ωi /ωpe
ωr /ωp,e
1.0
0.001
0.5 10-4
0.0 10-5
0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8
kλDe kλDe
Figure 11.3: ωr /ωpe (left) and −ωi /ωpe (right) as functions of kλDe for Langmuir waves in Maxwellian
electron plasma: red – numerical solution, blue – Eqs. (11.35) and (11.39).
.
where κ = kλDe . This leads to
r
ωi π −3 1 3
≈− κ exp − 2 − . (11.39)
|ωr | 8 2κ 2
Also note that in application to Langmuir waves, our assumption that ζe 1 can be expressed simply
as κ 1, because ω 2 ≈ ωpe 2
. In this case, κ−3 exp(−κ−2 /2) 1, so ωi ωr ; otherwise, our theory is
inapplicable. A comparison with the numerical solution of the dispersion relation is shown in Fig. 11.3.
Note that the fluid calculation of the Langmuir wave dispersion relation in Sec. 7 properly captured
ωr but not ωi . The kinetic calculation shows that in warm plasma, Langmuir waves have ωi < 0 and
therefore dissipate. This effect is known as Landau damping. It was predicted theoretically in Ref. [47],
and the first experimental observation was reported in Ref. [48]. The physical mechanism of Landau
damping will be discussed in Lecture 12.
χe = −
1
Z 0
(ζ ) ≈
1
+
i ζ e √π . (11.40)
e
2k 2 λ2D,e k 2 λ2D,e k 2 λ2D,e
108
LECTURE 11. ELECTROSTATIC WAVES IN ISOTROPIC MAXWELLIAN PLASMA 109
so one obtains
!−1
2
2 2 1 ωp,i k 2 λ2D,e k 2 Cs2
ωr = ωp,i 1+ 2 2 = = , (11.44)
k λD,e 1 + k 2 λ2D,e 1 + k 2 λ2D,e
This shows that ions also experience Landau damping. Unlike for Langmuir waves, this damping is
not exponentially small at small kλD,e ; rather, it is small due to the smallness of me /mi . The reason
for this different scaling will become clear after we discuss the physical mechanism of Landau damping
in the next lecture.
109
Lecture 12
In this lecture, we discuss the physical mechanism of Landau damping and the associated nonlinear
effects. Although our discussion is limited to electrostatic interactions in nonmagnetized plasma, the
qualitative effects to be considered are relevant also in more general settings.
with equilibria at kx̄ = 2πn. (Here, n is an integer, and we will assume es E0 < 0 and k > 0, but the
signs can always be changed by a variable transformation x̄ → x̄ + π/k.) Let us consider the vicinity
of an equilibrium with, say, n = 0. Close enough to the equilibrium, Eq. (12.1) can be approximated
with the equation of a harmonic oscillator,
¨ + Ω2b0 x̄ = 0,
x̄ (12.2)
110
LECTURE 12. LANDAU DAMPING AND KINETIC INSTABILITIES 111
ω
k
Figure 12.1: Trajectories of passing particles (non-shaded region) and trapped particles (shaded re-
gion). The arrows indicate the direction of the bounce motion. The dashed line corresponds to
v = ω/k.
(unbounded) trajectories from trapped (bounded) trajectories (Fig. 12.1). Passing trajectories, which
correspond to v0 > vt , extend from −∞ to +∞; particles on these trajectories have a nonzero average
velocity. In contrast, trapped trajectories, which have v0 < vt , are confined to a single wave period
and have zero average velocity in the moving frame, or the average velocity ω/k in the laboratory
frame.
The bounce frequency of trapped particles equals Ωb0 at v0 /vt 1 (even though Ωb0 is often called
“the” bounce frequency), but generally, it is found as follows. From Eq. (12.5), one obtains
where x∗ is the right stopping point. Let us adopt a new variable θ such that
This leads to
4K(r) . v2
Tb = , r = 02 , (12.10)
Ωb0 vt
where K is the complete elliptic integral of the first kind. The corresponding bounce frequency is
. 2π πΩb0
Ωb = = . (12.11)
Tb 2K(r)
111
LECTURE 12. LANDAU DAMPING AND KINETIC INSTABILITIES 112
Ωb /Ωb0
1.0
0.8
0.6
0.4
0.2
υ20 /υ2t
0.0 0.2 0.4 0.6 0.8 1.0
Figure 12.2: Ωb /Ωb0 vs. v02 /vt2 : solid blue – exact formula (12.11); dashed black – asymptotic (12.12).
As seen in Fig. 12.2, Ωb ∼ Ωb0 for all trapped particles except those very close to the separatrix, where
Ωb (r → 1) π
≈ → 0. (12.12)
Ωb0 ln[16/(1 − r)]
112
LECTURE 12. LANDAU DAMPING AND KINETIC INSTABILITIES 113
A A
(a) (b)
nonlinear stage
linear stage
nonlinear stage
linear stage
t t
Figure 12.3: A schematic of the evolution of the wave amplitude A(t) resulting from: (a) Landau
damping and (b) inverse Landau damping. The linear stage corresponds to A ∝ e ω t predicted by i
linear theory (Lecture 11). The dashed lines mark the asymptotic values at t → ∞.
f f
(a) (b)
υ υ
ω/k ω/k
Figure 12.4: Flattening of the distribution function near the resonance v = ω/k: dashed – initial
distribution, solid – distribution after saturation: (a) Landau damping – particles from the blue-
shaded region end up in the red-shaded region; (b) inverse Landau damping – particles from the
red-shaded region end up in the blue-shaded region.
These nonlinear effects limit the applicability of the linear Landau-damping theory presented in
Lecture 11. However, this theory can still be applicable at t & Ω−1 b0 if plasma is collisional. If the
collision rate νs exceeds Ωb0 , then the distribution function is kept close to Maxwellian at all times
and nonlinear flattening does not occur. That said, νs cannot be too large either. It should remain
small compared to kvT s , because otherwise collisional effects overshadow kinetic effects and the wave–
particle interaction ceases to be resonant, eliminating Landau damping. In summary then, for the
linear theory of collisionless Landau damping to apply at ωi . Ωb0 , one must have
113
LECTURE 12. LANDAU DAMPING AND KINETIC INSTABILITIES 114
Figure 12.5: Results of particle-in-cell simulations illustrating the development of the two-stream
instability (from Ref. [51]).
where we assumed ≈ 1 − ωp2 /ω 2 . Assuming also that trapped particles are accumulated at bottoms
of the wave troughs (this corresponds to a delta-shaped spatial distribution), one has
nt,k n̄t n̄t
∼− 2 =− . (12.16)
k 2 Uk k U |ee kE|
(The minus is due to the fact that the density has a maximum where the potential energy has a min-
imum, and n̄t denotes the spatial average of nt , i.e., the number of trapped particles per wavelength.)
This shows that the frequency shift is negative and is given by
δω 4πe2e n̄t ω2
∼− = − t2 , (12.17)
ωp |ee kE| Ωb
114
LECTURE 12. LANDAU DAMPING AND KINETIC INSTABILITIES 115
where ωt is the plasma frequency associated with the density n̄t ; i.e.,
s
. 4πe2e n̄t
ωt = . (12.18)
me
Also notice the unusual scaling |δω| ∼ E −1 . This scaling changes if trapped particles
√ are spread out
across the island. In fact, if the distribution is smooth, one can show that |δω| ∼ E, which is indeed
seen in laser–plasma interactions. For further details on dispersion and adiabatic dynamics of BGK-
like waves, see, for example, Refs. [53,54]. In particular, note that due to the nonlinear frequency shift
δω = δω(|E|), BGK-like waves with trapped particles can be subject to nonlinear instabilities [55–58].
115
Lecture 13
In this lecture, we extend our kinetic calculation of the plasma dielectric properties to three-dimensional
electromagnetic waves and magnetized plasma.
d t0 s ∂t
s
ms ∂v 0 0 0 0 0 [t ,x (t ),v (t )]
one then obtains a solution for f˜s at any given (t, x, v):
t
f˜s (t, x, v) = f˜s [t, x0 (t), v 0 (t)] = f˜(0)
s + d t0 Rs [t0 , x0 (t0 ), v0 (t0 )], (13.4)
t0
(0)
where the integral depends on x and v through Eqs. (13.3) and f˜s is an integration “constant”.
Using Eq. (8.19) for Rs , one obtains
t
v · k̂
f˜(i)
s = − d t 0 es n0s ∂f0s
ms ∂vc
δ cb 1 −
ω̂
+
vb k̂ c
ω̂
Ẽb . (13.5)
t0 [t0 ,x0 (t0 ),v 0 (t0 )]
116
LECTURE 13. DISPERSION PROPERTIES OF MAGNETIZED PLASMA 117
From here, one can find the induced current density (8.21) and infer the conductivity (8.22); then D E
is readily obtained.
117
Lecture 14
In this lecture, we apply the results of Lecture 13 to explore kinetic waves in magnetized isotropic
Maxwellian plasma without flows.
n2 In
−i n (In −
k⊥ nIn
An In0 ) An Bn
λ Ω λ
2
i k⊥ (I
i n (In − In ) An
0 n 0 0
Yn= In + 2λIn − 2λIn An n − In )Bn
, (14.2)
λ Ω
i k⊥ (I − I 0 )B
k⊥ nIn 2(ω − nΩ)
Bn − n n n In Bn
Ω λ Ω kk w 2
Z(ξn ) Z 0 (ξn ) 2 2
k⊥ w ω − nΩ
An = , Bn = − , λ= , ξn = , (14.3)
kk w 2kk 2Ω2 kk w
Exercise 14.1: Identify the limit in which Eqs. (14.1)–(14.3) reproduce the cold-plasma dielec-
tric tensor (6.13).
As usual, the dispersion relation and the equation for the field polarization are
118
LECTURE 14. WAVES IN MAGNETIZED PLASMA 119
xx − Nk2
xy xz + N⊥ Nk
DE = yx yy − N 2 yz . (14.5)
2
zx + N⊥ Nk zy zz − N⊥
Below, we will focus on waves propagating perpendicularly to the magnetic field (Nk → 0). For
parallel propagation (N⊥ → 0), see Problem PIV.7.
so there are two types of modes. Ordinary (O) modes, which are polarized along the z axis, satisfy
zz − N 2 = 0 (14.9)
and subsume the cold O wave as a special case. Extraordinary (X) modes, which are polarized in the
plane transverse to the z axis, satisfy
xx xy
det =0 (14.10)
yx yy − N 2
and subsume the cold X wave as a special case. The effects introduced by kinetic corrections for
O and X waves are similar, but Eq. (14.10) is somewhat richer than Eq. (14.9), so in this lecture we
will focus on X waves. For a more comprehensive discussion, see Refs. [60, 61].
119
LECTURE 14. WAVES IN MAGNETIZED PLASMA 120
Exercise 14.2: For an x-polarized wave, the y component of Eq. (14.11) is yx (ω, k) = 0, which
is inconsistent with Eq. (14.12). Explain why Eq. (14.12) should be preferred over this equation.
For clarity, let us focus on electron frequencies (of order ωuh ∼ Ωe or higher), where the ion
contribution is negligible. Then, = 1 + χe ,
2 ∞ 2
ωpe
e −λe n In (λe )
X
(χe )xx ≈ An
ω n=−∞ λe
2 ∞ 2
ωpe
e −λe n In (λe ) 1
X
≈−
ω n=−∞ λe ω − nΩe
2 X∞ 2
ωpe
=−
ω n=1
e −λe n In (λe )
λe
1
ω − nΩe
+
1
ω + nΩe
2 X∞ 2
ωpe
=−
ω n=1
e −λe n In (λe )
λe
2ω
ω 2 − (nΩe )2
2 X∞
ωpe n2
=− 2
Ωe n=1
e −λe 2In (λe )
λe (ω/Ωe )2 − n2
, (14.13)
with αn and H illustrated in Fig. 14.1. The modes described by Eq. (14.15) are known as (electrostatic)
electron Bernstein waves (EBW). Their frequencies can be approximately calculated as follows. Note
that
n−1 r
λ 2 −3/2
αn (λ → 0) ∼ , αn (λ → ∞) ∼ λ , (14.16)
2 π
so there is always a mode with ω(λe → 0) → ωuh . Also, clearly, there are modes with ω → n|Ωe | at
both λe → 0 and λe → ∞. To connect these limits, let us consider EBW at β 1 (very underdense
plasma). In this case, H(q, λe ) has to be large to satisfy Eq. (14.15), so q must be close to some
integer n. Then, one can retain only the nth term in the sum in Eq. (14.15), so the latter becomes
n2 αn (λe )
1 − β2 ≈ 0. (14.17)
q 2 − n2
ω 2 = n2 [Ω2e + ωpe
2
αn (λe )]. (14.18)
120
LECTURE 14. WAVES IN MAGNETIZED PLASMA 121
α1 α2,3,4
1.0 (a) 0.10 (b)
0.8 0.08
0.6 0.06
0.4 0.04
0.2 0.02
λ λ
1 2 3 4 5 1 2 3 4 5
H
3
(c)
2
q
0.5 1.0 1.5 2.0 2.5 3.0 3.5
-1
-2
-3
Figure 14.1: Upper row: αn (λ) for various n. Lower row: H(q, λ) at λ = 0.2.
while at n > 1, one has ω → n|Ωe | at both λe → 0 and λe → ∞. The situation is somewhat different
in denser plasmas, as illustrated in Fig. 14.2.
Notably, all solutions of Eq. (14.15) are real. This means EBW as considered here do not dissipate.
The absence of dissipation is due to the fact that these waves propagate transversely to B 0 and
cannot resonantly interact with plasma particles. However, this ceases to be the case when relativistic
effects are taken into account [61]. Also, kk cannot be strictly zero in practice, and the ideal-plasma
approximation is inapplicable at very large k⊥ . In real plasmas, EBW dissipate efficiently, because they
have a small group velocity and thus take a long time to propagate through a plasma of a given length.
This makes these waves convenient vehicles for depositing energy into plasma, especially because EBW
can propagate in dense plasmas which electromagnetic waves cannot penetrate (Sec. 14.3.3). However,
note that EBW cannot be launched by antennas placed outside the plasma, because EBW cannot
propagate outside plasma. To understand how to launch these waves, electromagnetic effects have to
be considered. Qualitatively, these effects are understood as follows.
121
LECTURE 14. WAVES IN MAGNETIZED PLASMA 122
ω/Ωe ω/Ωe
5 (a) 5 (b)
4 4
3 3
2 2
1 1
λe λe
0.5 1.0 1.5 2.0 0.5 1.0 1.5 2.0
2 2
Figure 14.2: The frequencies ω of the first four EBW in overdense plasma vs. λe = k⊥ vT e /Ω2e : left
– β = 3 (ωuh ≈ 3.16|Ωe |); right – β = 4 (ωuh ≈ 4.12|Ωe |). The dashed lines correspond to ω = n|Ωe |
with integer n and also to ω = ωuh .
• The dispersion curve of the cold-plasma X wave crosses the resonances at ω = n|Ωe |, so with DK
modes taken into account, one can expect the new dispersion curves to exhibit avoided crossing
(Sec. 6.4.5) near the resonances. In other words, the X wave should continuously transform into
DK modes.
• Homogeneous Maxwellian plasma has no free energy, so the corresponding waves are stable. By
general theory of avoided crossing (Sec. 6.4.5), resonant coupling of two stable waves modifies
the corresponding dispersion curves that there remain two real frequencies for each real k.
These considerations are sufficient to plot the corresponding dispersion curves unambiguously. This
is illustrated in Fig. 14.3 for n = 1 in underdense plasma. There is no commonly accepted names for
the individual branches, but one can identify them locally based on the closest asymptotic. Then,
one can say that the electrostatic EBW continuously transforms into the cold X wave, and the latter
continuously transforms into the DK mode.
122
LECTURE 14. WAVES IN MAGNETIZED PLASMA 123
ω/Ωe N2
1.15 (a) 5 (b)
1.10 4
1.05 3
1.00 2
0.95 1
λ ω/Ωe
0.2 0.4 0.6 0.8 1.0 0.6 0.8 1.0 1.2
Figure 14.3: Electromagnetic dispersion curves for β = 0.6, vT2 e /c2 = 0.1, and ω in the vicinity of ωuh :
(a) – in the (λ, ω/|Ωe |) space; (b) – same in the (ω/|Ωe |, N 2 ) space. On the left: the blue dashed
curve corresponds to the electrostatic approximation (14.15) for the EBW with n = 1; the red dashed
line corresponds to ω = |Ωe |. On both figures: the black dashed curves correspond to the cold-plasma
X wave.
To deposit power into the core then, one solution is to launch an X wave from the high-B0 side and
bounce it off the upper-hybrid resonance (S = 0 curve in Fig. 14.4). When the X wave approaches the
upper-hybrid resonance, the wave number k continues to grow (Exercise 14.3). Then, λe eventually
shifts to the right from the local maximum in the dispersion curve [Fig. 14.3(a)] and gradually turns
into the electrostatic EBW. The group velocity changes its sign then,
2 2
∂ω ∂λe ∂ω k w ∂ω
k · vg = k · = k· = 2
< 0, (14.20)
∂k ∂k ∂λe Ω ∂λe
so the wave starts to propagate backwards. (This is called X–B conversion.) Now that the wave
energy is in the EBW, it is insensitive to the cutoff at L = 0, so it can penetrate the dense plasma
core, as desired.
Exercise 14.3: Using ray equations, explain why k continues to grow as the X wave is trans-
forming into an EBW.
An alternative solution is to launch an O wave at a special angle that ensures complete O–X
mode conversion at the location where ωp2 = ω 2 (Fig. 14.5). Then, the X wave propagates to the
upper-hybrid resonance and transforms into the EBW as usual. (This is called O–X–B conversion.)
123
LECTURE 14. WAVES IN MAGNETIZED PLASMA 124
Figure 14.4: The CMA diagram for cold-plasma waves (copied from Ref. [1]). The X wave cannot
propagate in the area sandwiched between the curves R = 0 and S = 0, and it also cannot propagate
to the left from the curve L = 0.
124
LECTURE 14. WAVES IN MAGNETIZED PLASMA 125
N⟂ N⟂ N⟂
2 (a) N||2 < N
2 2 (b) N||2 = N
2 2 (c) N||2 > N
2
1 1 1
-1 -1 -1
-2 -2 -2
Figure 14.5: The transverse refraction index for cold-plasma waves in a homogeneous magnetic field
vs. ωp2 /ω 2 . The three figures show different scenarios depending on how the initial Nk2 relates to the
optimum value N̄k2 = |Ωe |/(ω + |Ωe |). The black points mark cutoffs, and the vertical dashed lines
mark the location of the upper-hybrid resonance. See Problem PIII.2 for details.
relation is now
2 ∞ 2
ωpi
e −λi n In (λi )
X
0 = xx = 1 + (χe )xx + An (ζi ). (14.21)
ω n=−∞ λi
Because
λe w 2 Ω2 me
∼ e2 i2 ∼ 1, (14.22)
λi w i Ωe mi
2
one can treat electrons as cold for all λi . mi /me , so (χe )xx ≈ ωpe /Ω2e [Eq. (14.13)]. Then, Eq. (14.21)
acquires the same form, up to coefficients, as the corresponding equation for electrostatic EBW:2
2 ∞
ωpi n2 In (λi )
e −λ
X
1+ 2 /Ω2 ω
i
An (ζi ) = 0. (14.23)
1 + ωpe e n=−∞
λi
This is impractical, so nonzero Nk is used instead, in which case the upper-hybrid resonance can be
made accessible from vacuum (almost). For further details, see Ref. [63].
2 If
ω Ωi and λi 1, many terms in the sum (14.23) may be nonnegligible. Then, it may be more convenient to
use the expression that does not involve infinite series [59]. Notwithstanding large λi , the ion susceptibility may then
still be approximated well with the corresponding cold-plasma formula. For example, see Appendix B in Ref. [62].
125
Lecture 15
Collisionless dissipation in
magnetized plasma
In this lecture, we discuss mechanisms of collisionless dissipation of the wave power in magnetized
isotropic Maxwellian plasma without flows.
In the presence of a strong magnetic field B 0 = ēz B0 , the particle trajectory can be locally expressed as
where vk is the average velocity parallel to B 0 , and x̃⊥ is the transverse quiver displacement of the
particle from a field line that the particle follows on average, hx̃⊥ i = 0. The constant x0 can always
be made zero by redefining the origin. Then, assuming the notation
P = Re g(t) e −i (ω−kk vk )t ,
D E
(15.4)
where g is given by
⊥ (t) · E ⊥ + vk E z e
. ˙ i k ·x̃ (t)
g(t) = es x̃ ⊥ ⊥
. (15.5)
Because g is periodic in time with period 2π/Ωs , where Ωs is particle’s gyrofrequency, this function
can be represented as a Fourier series
∞
gn e i nΩs t
X
g(t) = (15.6)
n=−∞
126
LECTURE 15. COLLISIONLESS DISSIPATION IN MAGNETIZED PLASMA 127
so substantial energy exchange is possible when there exist an integer n such that
ω ≈ kk vk + nΩs . (15.8)
This is called resonant absorption at the nth harmonic.
Different particles have different vk ; however, assuming that the distribution of vk is centered
around zero (i.e., there are no average flows) and Ωs are much larger than kk ws , plasma as a whole
can effectively absorb the wave power only if
ω ≈ nΩs (15.9)
for some of its species s. Below, we show how this condition also flows from general kinetic theory,
and we also discuss how to explicitly calculate the wave absorption in Maxwellian plasma using results
from previous lectures.
127
LECTURE 15. COLLISIONLESS DISSIPATION IN MAGNETIZED PLASMA 128
√
1 d √ −ξn2
πe ξn e −ξn = (ξn w) (Im An ).
π 2
Im Bn = − = (15.15b)
2kk d ξn kk
This leads to
√ ∞
!
π ωp2 X −ξn2 −λ
e e Ȳ n
X
A = (χs )A , (χs )A = , (15.16)
s
kk w ω n=−∞
s
ω|E z |2 ωp2
h(ξ0 ) = ξ03 e −ξ0 .
. 2
PLD,s = √ h(ξ0 ) , (15.20)
4 π ω2 s
1/2
In a cold plasma, ξ0 is large, so h(ξ0 ) is exponentially small; then because ξs ∝ ws−1 ∝ ms , the
1/2
electron contribution dominates. In hot plasma, ξ0 is small, so h(ξ0 ) ≈ ξ03 , and PLD,s ∝ ms ; in this
case, the ion contribution dominates.
128
LECTURE 15. COLLISIONLESS DISSIPATION IN MAGNETIZED PLASMA 129
kÞ = 0 kÞ ¹ 0
Figure 15.1: A schematic of magnetic field lines in magnetized plasma perturbed by an electromagnetic
wave: left k⊥ = 0, right – k⊥ 6= 0. The x axis is vertical, and the z axis is horizontal.
.
produce oscillations of the magnitude of the total field B = |B 0 + B̃| (Fig. 15.1):
q
B̃ = B̃⊥ 2 + (B + B̃)2 − B
0 0
q
≈ B02 + 2B 0 · B̃ − B0
c c
≈ B̃z = (k × Ẽ)z = k⊥ Ẽy . (15.21)
ω ω
For simplicity, suppose that ω Ωs . Then, the oscillations of B produce an oscillating diamagnetic
2
force F̃ ≈ −µ ∂z B̃ on each particle along the z axis, where µ ≈ ms v⊥ /2B0 is particle’s magnetic
moment. In the complex form, one can write the ensemble-average of this force as
which can be viewed as an effective longitudinal electric field, Ẽeff,s = hF̃ is /es . Like the true longitu-
dinal electric field, Ẽeff,s produces Landau damping, which is known as transit-time magnetic pumping
(TTPM). Accordingly,
PTTMP,s ∼ PLD,eff,s
!
ω|Ẽeff,s |2 ωp2 3 −ξ2
= √ ξ e 0
4 π ω2 0
s
!
ω|Ẽy |2 c2 T 2 ωp2 ω 2 ω
= √ kk2 k⊥
4 π
2
e2 B02 ω 2 ω 2 kk2 w2 kk w
e −ξ02
s
!
2 2 ω2
ω|Ẽy | T 2 k⊥
= √
4 π
p ω
m 2 w 2 Ω2 ω 2 k k w
e −ξ0
2
s
!
ω|Ẽy |2 ωp2
λ 2 ξ0 e −ξ0
2
= √ . (15.23)
8 π ω
s
129
LECTURE 15. COLLISIONLESS DISSIPATION IN MAGNETIZED PLASMA 130
ω|B̃z |2
√ (βξ0 e −ξ0 )s .
2
= (15.24)
8 π
√
The maximum of this function is achieved at ξ0 ∼ 1, so PTTMP, max ∼ βω|B̃z |2 /(8 π).
A more rigorous expression for the TTMP comes from the general formula (15.10), where it
corresponds to the contribution of E y and n = 0:
ω
PTTMP = (A,yy )n=0 |E y |2 , (15.25)
8π
which leads to a result consistent with Eq. (15.23). That said, generally, one should also take into
account the “cross terms” Pyz,n=0 and Pzy,n=0 , which can be of the same order as Pyy,n=0 and
Pzz,n=0 [64]. In this sense, Landau damping and TTMP are not always separable.
Note that the coefficients decrease rapidly with n, so the absorption at high-order resonances is
relatively weak. Also, efficient heating is possible only when (ξ±n )s are small enough, i.e., ω = ±nΩs ,
and the field is polarized such that its corresponding component Ẽ± = Ẽx ± i Ẽy does not vanish.
.
Like Landau damping, cyclotron heating can saturate through nonlinear effects. To understand
these effects qualitatively, let us consider a simple model in which a transverse wave propagates strictly
parallel to B 0 , i.e., k = kk ēz . The wave electric field Ẽ 0 in the frame moving at the phase velocity
v p = (ω/kk )ēz can be expressed through the wave fields in the laboratory frame using the well known
relativistic transformation [9]. Specifically, one finds that
vp
γp−1 Ẽ 0 = Ẽ + × B̃
c h
ω c i
= Ẽ + ēz × k × Ẽ
ck ω
= Ẽ + ēz × (ēz × Ẽ)
= Ẽ + ēz (ēz · Ẽ) − Ẽ(ēz · ēz )
= 0, (15.27)
where γp is the Lorentz factor associated with v p . Then, the particle energy in this frame is conserved,
0 2
(v⊥ ) + (vk0 )2 = const. (15.28)
(From now on, we assume that both v 0 and vp are nonrelativistic.) In the laboratory-frame variables,
this can be expressed as follows:
This describes a (semi-)circle in the (vk , v⊥ ) plane shifted by ω/kk along vk axis (Fig. 15.2). Depending
on the initial phase at which a particle enters the wave, the particle may gain or lose energy from the
wave and thus can move up or down such a circle. This causes diffusion of the particle distribution
130
LECTURE 15. COLLISIONLESS DISSIPATION IN MAGNETIZED PLASMA 131
Figure 15.2: A schematic of the diffusion paths (sold curves) given by Eq. (15.29). The vertical red
line denotes the resonance (15.30). Also shown are isosurfaces of the Maxwellian distribution (dashed
curves). At sufficiently high ω/(kk ws ), the wave heats particles with large vk2 , and the majority of
2
those have low v⊥ (because the distribution function falls off as ∼ exp[−(vk2 + v⊥
2
)/ws2 ]), so on average,
particles are heated up. The shaded region corresponds to the trapping area.
in the velocity space along the circles (15.29), which are hence called diffusion paths. The diffusion
coefficient depends on the distance from the resonance (15.8) and is maximized at
To better understand how individual particles interact with the wave, let us differentiate Eq. (15.29)
with respect to time; then, one obtains
d vk 1 d v⊥2
=−
dt vk − ω/kk d t 2
=
kk
v⊥ ·
d v⊥
ω(1 − kvk /ω) dt
kk es 1 1
= v⊥ · Ẽ + v × B̃ + v × B 0 . (15.31)
ω(1 − kvk /ω) ms c c ⊥
Notice that
1 1
Ẽ + v × B̃ = Ẽ + v × (k × Ẽ)
c ⊥ ω ⊥
k(v · Ẽ) (k · v)
= Ẽ + − Ẽ
ω ω ⊥
kvk (v · Ẽ)
= Ẽ 1 − + k⊥
ω ω
k k vk
= Ẽ 1 − (15.32)
ω
v ⊥ · (v × B 0 )⊥ = v ⊥ · (v ⊥ × B 0 )⊥ = (v ⊥ × v ⊥ ) · B 0 = 0. (15.33)
131
LECTURE 15. COLLISIONLESS DISSIPATION IN MAGNETIZED PLASMA 132
Hence, we obtain
d vk kk es
kvk
es kk v ⊥
= v⊥ · Ẽ 1 − = · Ẽ. (15.34)
dt ω(1 − kk vk /ω) ms ω ms ω
Then,
where we used that vx ∓ i vy ∝ e ±i Ωs t (cf. Sec. 6.1). Substituting this into Eq. (15.34) leads to
d 2z0 =
es (eff)
E sin(kk z 0 ), (15.38)
dt ms 0
where
. ω ∓ Ωs (eff) . k k v⊥
z0 = z − t + const, E0 = |E |, (15.39)
kk ω
and kk > 0 will be assumed for clarity. This is similar to Eq. (12.1) that we used to explore the
nonlinear saturation of Landau damping earlier; the only difference is that the resonance is now
(eff)
shifted by Ωs and E0 is replaced with E0 . This means that the nonlinear saturation mechanism for
cyclotron damping is similar to that of Landau damping, except that the trapping width (Fig. 15.2)
(eff)
is now of order (es E0 /ms kk )1/2 and the corresponding bounce time is
s s
ms ms ω
T ∼ (eff)
= . (15.40)
|es E |kk |es E0 |kk kk v⊥
0
If v⊥ is small, reaching the nonlinear stage through cyclotron damping takes longer than at Landau
damping (for given kk es E0 /ms ). This is understood from the fact that a resonant particle with small
v⊥ travels primarily upward in Fig. 15.2, so it takes longer for vk to become nonresonant.
This picture can be significantly modified, though, when waves have a broad spectrum. A particle
interacting with multiple waves that have different phase velocities is not necessarily constrained to
a single diffusion path but can diffuse in wider regions of the velocity space. Then, the nonlinear
saturation of the wave absorption is governed by a very different equations. This regime will be
discussed, within a simpler model, in Lecture 16.
132
Lecture 16
Quasilinear theory
In this lecture, we discuss the nonlinear evolution of the particle distribution driven by self-consistent
waves with a sufficiently wide and dense spectrum. For simplicity, one-dimensional nonmagnetized
electron plasma will be assumed.
16.1 Introduction
16.1.1 One wave: nonlinearities due to trapped particles
The linear-wave approximation used in previous lectures implies
As discussed in Lecture 12, this approximation holds in collisionless plasma only on time scales t .
Ω−1
b0 , where Ωb0 ∝ Ẽ
1/2
is the bounce frequency. At t & Ω−1 b0 , the distribution function acquires
structure with the characteristic velocity scale equal to the trapping-island size vt ∼ Ωb0 /k. Then,
near the resonance, ∂v fs scales as fs /vt ∼ Ẽ −1/2 . This makes the third term in the Vlasov equation
which is known as the Chirikov criterion. If the Chirikov criterion is not satisfied (vt1 + vt1
|vp1 − vp2 |), then any given particle can be resonant to only one of the waves. In this case, each wave
133
LECTURE 16. QUASILINEAR THEORY 134
υ υ
(a) (b)
υ p1 υ p1
υ p2
υ p2
x x
Figure 16.1: A schematic of the trapping islands produced by two waves with different phase velocities,
vp1 and vp2 : (a) the Chirikov criterion is not satisfied (shaded are individual trapping islands); (b) the
Chirikov criterion is satisfied (shaded is the region of stochastic motion).
saturates more or less independently from the other one, i.e., as described in Lecture 12. But if the
Chirikov criterion is satisfied (vt1 + vt1 . |vp1 − vp2 |), then the islands overlap and trapped particles
no longer “belong” to a particular wave but are rather “shared” by the two waves. The trajectories
of such particles are generally stochastic and extend to the whole interval vp2 − vt2 . v . vp1 + vt1 .
Then, clearly, if the two waves saturate, they saturate together.
are allowed, so the distribution of vp is continuous; then, the Chirikov criterion is automatically satisfied even at
vanishingly small amplitudes. In contrast, if the boundary conditions are periodic (as, for example, in a tokamak for
toroidal and poloidal modes), then vp are quantized and thus satisfying the Chirikov criterion requires finite amplitudes.
134
LECTURE 16. QUASILINEAR THEORY 135
∂f0 ∂ f˜ ∂ f˜ ee ∂f0 ee ∂ f˜
+ +v + Ẽ + Ẽ = 0. (16.5)
∂t ∂t ∂x me ∂v me ∂v
By performing spatial averaging, one obtains the following equation for f0 :
* +
∂f0 ee ∂ f˜
+ Ẽ = 0. (16.6)
∂t me ∂v
The equation for f˜ is obtained by subtracting the equation for f0 from the original equation:
* +
∂ f˜ ∂ f˜ ee ∂f0 ee ∂ f˜ ee ∂ f˜
+v + Ẽ + Ẽ − Ẽ = 0. (16.7)
∂t ∂x me ∂v me ∂v me ∂v
| {z }
N
When the dynamics is regular, a small field can, in principle, make f˜ order-one over large enough
time. But here, we assume that the particle dynamics is stochastic, so correlations die out fast; then f˜
remains small at all times and the term N , which is nonlinear yet has zero average, can be omitted.
This is called the quasilinear approximation, and it leads to
∂ f˜ ∂ f˜ ee ∂f0
+v + Ẽ ≈ 0. (16.8)
∂t ∂x me ∂v
By applying the spatial Fourier transform to this equation, we also obtain
∂ f˜k (t, x, v)
+ i kv f˜k (t, x, v) +
ee ∂f0 (t, v)
Ẽk (t) = 0. (16.9)
∂t me ∂v
As discussed earlier (Sec. 8.3), the general solution of this equation can be written as
t 0
f˜k (t, v) = f˜k (t0 , v) e −i kvt − e −i kvt d t0 e i kvt Ẽk (t0 ) ∂f0∂v
ee 0 (t , v)
. (16.10)
me t0
The function f0 (t0 , v) in the integrand is slow, whereas e i kvt and Ẽk (t0 ) are rapidly oscillating. Let
0
where Ẽk,0 is a constant, and θk is a complex phase, so one can introduce a local complex frequency
ωk (t) = −dt θk (t) = −dt arg[Ẽk (t)] + idt ln |Ẽk (t)| ≡ ωk,r + i ωk,i .
.
(16.12)
135
LECTURE 16. QUASILINEAR THEORY 136
d τ iFϑ0(τ(τ)) d ed τ
t i ϑ(τ )
J (t) =
t0
t t
d F (τ )
i ϑ (τ ) e d e
F (τ ) i ϑ(τ )
= 0 − τ i ϑ(τ )
t0 t0 d| τ i{zϑ0 (τ ) }
O(εF )
∗
Thus, Ẽ−k = Ẽk . This means that and ω−k = −ωk∗ = −ωk,r + i ωk,i , so
Because ωk and f0 are assumed slow compared to e i θk (t) , one can represent f˜k as2 (Box 16.1)
Assuming ωk,i > 0, the second term eventually dominates, so one obtains
The actual distribution f˜(t, x, v) is obtained by taking the inverse Fourier transform of f˜k (t, v):
f˜(t, x, v) =
1 ∞
d k f˜k (t, v) e i kx = − mee
∞
dk Ẽk (t) e i kx ∂f0 (t, v)
. (16.17)
2π −∞ e −∞ 2π i [kv − ωk (t)] ∂v
Let us substitute this and the expression for Ẽ(t, x) into the equation for f0 :
* +
∂f0 ee ∂ f˜
=− Ẽ
∂t me ∂v
∞
d k Ẽk (t)e i kx ∂f0 (t, v)
* +
∂ e2
= Ẽ(t, x)
∂v m2e −∞ 2π i [kv − ωk (t)] ∂v
*
d d k Ẽk (t)e i kx
+ !
∞ ∞
e2 k0 ∗
Ẽk0 (t)e
∂ −i k 0 x ∂f0 (t, v)
=
∂v m2e −∞ 2π −∞ 2π i [kv − ωk (t)] ∂v
d k 0 ∞ d k Ẽk∗0 (t)Ẽk (t) D i (k−k0 )x E ∂f0 (t, v)
!
∞
e2
=
∂
∂v m2e −∞ 2π −∞ 2π i [kv − ωk (t)]
e ∂v
∞
d k 0 ∞ d k Ẽk∗0 (t)Ẽk (t) 2π
!
∂ e2 0 ∂f0 (t, v)
= δ(k − k )
∂v m2e −∞ 2π −∞ 2π i [kv − ωk (t)] L ∂v
2 Although Eq. (16.15) is correct as the leading-order approximation, retaining corrections O(∂t f0 ) and O(∂t ωk ) is
generally necessary to keep the quasilinear model truly conservative. For Langmuir turbulence, this is not a big issue
somewhat accidentally. See the footnotes below, also see Lecture 17.
136
LECTURE 16. QUASILINEAR THEORY 137
dk
!
∞
∂ e2 |Ẽk (t)|2 ∂f0 (t, v)
= . (16.18)
∂v m2e −∞ 2πL i [kv − ωk (t)] ∂v
and Uk is understood as the spectral density of the electric-field energy per unit volume, because
∞ ∞
d k0 ∗ d
2
Ẽk0 (t)e Ẽk (t)e
Ẽ 1 −i k 0 x k i kx
=
8π 8π −∞ 2π −∞ 2π
∞
=
1 d k0 ∞ d k Ẽ ∗ (t)Ẽ (t) 2π δ(k − k0 )
0 k
8π −∞ 2π −∞ 2π k L
∞
=
1 d k |Ẽ (t)|2
k
8π −∞ 2πL
∞
= d k Uk . (16.23)
−∞
Although Eq. (16.22) contains imaginary unit, the corresponding D is real, as easily seen from the
fact that Uk = U−k and ω−k = −ωk∗ . Furthermore, to the extent that the nonzero value of ωk,i is
negligible, one readily obtains, using Landau’s rule, that
. 8πe2
D(t, v) ≈ D̄(t, v) = Im d k kv − Uωk (t)
m2e L k (t) − i 0
2 2 ∞
δ(kv − ωk,r ) Uk d k.
16π e
= (16.24)
m2e 0
1 1 d |Ẽk,0 e i θk (t) |2
=
2πL 8π dt
1 |Ẽk,0 e i Re θk (t) |2 de −2Im θk (t)
=
2πL 8π dt
1 |Ẽk,0 e i θk (t) 2
|
= 2ωk,i , (16.25)
2πL 8π
or in other words,3
d Uk = 2ωk,i Uk . (16.26)
dt
Since f˜ that we derived earlier is governed by the same equation as in stationary plasma, the frequen-
cies ωk (t) can be found using the GO dispersion relation:
4πe2
1−
me k 2
d v v∂−v fω0 (t,(t)/k
v)
= 0. (16.27)
L k
where D is given by Eq. (16.22) and ωk is found from Eq. (16.27). These equations conserve4 the total
number of particles:
d ∞ d v f (t, v) = ∞ d v ∂ D ∂f0 = 0,
(16.29)
d t −∞ 0
−∞ ∂v ∂v
the total momentum:
d ∞ d v m vf (t, v) = ∞ d v m v ∂ D ∂f0
d t −∞ e 0
−∞
e
∂v ∂v
∞
=− d v me D ∂f
∂v
0
−∞
∞ 2
=− d v d k i (kv2U−kω ) 4πe
me
∂f0
∂v
−∞ L k
∞ 2
= 2i d k Uk 4πe d v kv∂v−f0ω
−∞ me L k
∞
4πe2
= 2i d k kUk m 2
d v v −∂vωf0/k
−∞ ek L k
∞
= 2i d k kUk = 0 (16.30)
−∞
3 Strictly speaking, Eq. (16.26) should be the equation for the mode action I rather than for U . For Langmuir
k k
waves in homogeneous plasma, Ik ∝ Uk /ωk and ωk ≈ ωp = const, so the equations for Ik and Uk are usually not
distinguished. A more rigorous approach is described in Lecture 17.
4 As discussed in footnotes 2 and 3, the standard quasilinear theory presented here is oversimplified in two aspects.
138
LECTURE 16. QUASILINEAR THEORY 139
(where we substituted the dispersion relation and used the fact that Uk = U−k ), and the total energy:
d ∞ d v me v2 f (t, v) + ∞ d k U (t)
d t −∞ 2
0
−∞
k
∞ 2 ∞
= d v me2v ∂t f0 + d k U̇k
−∞ −∞
∞ 2
∞
= d v me2v ∂v∂
D
∂f0
∂v
+ d k 2ωk,i Uk
−∞ −∞
∞ ∞
=− d v me vD ∂f
∂v
0
+ d k 2ωk,i Uk
−∞ −∞
∞ 2 ∞
=− d k d v i (kv2U−k ω ) 4πe v ∂f0
me ∂v
+ d k 2ωk,i Uk
−∞ L k −∞
∞
= 2i d k (ωk − i ωk,i )Uk
−∞
∞
= 2i d k ωk,r Uk
−∞
= 0. (16.31)
d d v f (t, v) = 0, (16.33a)
dt 0
d
d t d v me vf0 (t, v) = 0,
(16.33b)
d 2 ∞
Because f0 evolves due to Uk , this theory is not entirely linear. However, since the field spectrum is
determined by the local linear dispersion relation, this theory is called quasilinear.
139
LECTURE 16. QUASILINEAR THEORY 140
f f
(a) (b)
υ υ
Figure 16.2: Quasilinear evolution of the distribution with a bump on tail: (a) initial distribution;
(b) saturated distribution with a quasilinear plateau. The areas between the dashed curve and the
solid curve in (b) are equal as required by particle conservation.
Note that the formation of a quasilinear plateau indicates that the average momentum of the
resonant particles decreases. But the electrostatic field carries no momentum, so this momentum
change must be absorbed by the background distribution, which apparently experiences adiabatic
transformation while tail particles interact with the waves resonantly. This transformation can be
explained ad hoc by the difference between D and D̄ (16.24) [1]. A more systematic approach is
presented in Lecture 17.
140
Lecture 17
In this lecture, we generalize quasilinear theory presented in Lecture 16. In particular, we introduce
the concept of dressed particles (also called oscillation centers) and ponderomotive effects.
141
LECTURE 17. DRESSED PARTICLES AND PONDEROMOTIVE FORCES 142
.
D̂ab = e2 hÊ a ĜÊ b i . (17.4)
It is straightforward to show that the Weyl symbol of D̂ab can be expressed through the Weyl symbol
of Ĝ as follows:
D(t, x, ω, k) = e2 d ω0 d k0 G(ω − ω0 , k − k0 ) hW E i(t, x, ω0 , k0 ). (17.5)
Here, hW E i is the average Wigner tensor of the electric field, i.e., the spectrum of the symmetrized
.
two-point correlation matrix Cab (t, x, τ, s) = hẼa (t + τ /2, x + s/2) Ẽb (t − τ /2, x − s/2)i:
hWE,ab i(t, x, ω, k) =
1
(2π)n+1
d τ d s Cab (t, x, τ, s) e −i k·s+i ωτ , (17.6)
.
which is also understood as the average Weyl symbol of Ŵ E = (2π)−(n+1) |Ẽi hẼ|. (Here, n is
the number of spatial dimensions.) Because the field is electrostatic, it is convenient to express
hW E i through the average Wigner function of the electrostatic potential, i.e., the spectrum of the
.
symmetrized two-point correlation function Cϕ (t, x, τ, s) = hϕ̃(t + τ /2, x + s/2) ϕ̃(t − τ /2, x − s/2)i:
hWϕ i(t, x, ω, k) =
1
(2π)n+1
d τ d s Cϕ (t, x, τ, s) e −i k·s+i ωτ , (17.7)
.
which is the average Weyl symbol of Ŵ ϕ = (2π)−(n+1) |ϕ̃i hϕ̃|. Because
Ŵ E = k̂ Ŵ ϕ k̂, (17.8)
i k ∂hWϕ i − k ∂hWϕ i ,
hWE,ab i ≈ ka kb hWϕ i − b a (17.9)
2 ∂xa ∂xb
assuming the averaged quantities are smooth enough such that GO approximation is applicable.1
Similarly, because f0 is only weakly inhomogeneous, one can Weyl-expand D̂ like in Lecture 3, that
is, by using
and then applying the inverse Wigner–Weyl transform. After a tedious but straightforward calculation,
one arrives at the following diffusion equation:
∂F ∂H ∂F ∂H ∂F ∂ ∂F
+ · − · = D̄ab , (17.11)
∂t ∂p ∂x ∂x ∂p ∂pa ∂pb
D̄ab = πe 2
d ω d k ka kb hWϕ i(t, x, k · v, k), (17.12)
where nonlinearities beyond that of the second order in Ẽ have been neglected. (Note that Wϕ is
evaluated at ω = k · v, which indicates that diffusion is governed by resonant particles.) Here,
p2
1 ∂ ∂f0
F = f0 + Vab , H= + Φ. (17.13)
2 ∂pa ∂pb 2me
1 Strictly speaking, W can be delta-shaped [Eq. (17.17)], but we will be interested only in integrals of W . Con-
ϕ ϕ
tributions of higher-order derivatives of Wϕ to those integrals is small even for delta-shaped Wϕ when J and ωr are
smooth functions.
142
LECTURE 17. DRESSED PARTICLES AND PONDEROMOTIVE FORCES 143
The function H serves as an effective OC Hamiltonian, and Φ is known as the ponderomotive potential,
or more precisely, the ponderomotive energy of an OC. Specifically,
e2 ka kb hWϕ i(t, x, ω, k)
Vab =
∂
∂ϑ
d ω dk
ω−k·v+ϑ
, (17.14a)
ϑ=0
2
khWϕ i(t, x, ω, k)
Φ=
∂
∂p
dω dk e 2me (ω − k · v)
. (17.14b)
.
Also note that the OC velocity is not quite v = p/me but is rather given by
. ∂H p ∂Φ ∂Φ
V = = + ≡v+ . (17.15)
∂p me ∂p ∂p
(Because V − v = O(Ẽ 2 ), the velocities V and v are interchangeable in Eqs. (17.12) and (17.14)
within the accuracy of this theory.) Similar equations can be derived for plasma interaction with
electromagnetic waves, except the formulas for D̄ab , Vab , and Φ will be different.
e2 k 2 |ϕ̃c |2
Φ= , (17.16)
4me (k · v − ωr )2
where ϕ̃c is the complex amplitude of the actual potential ϕ̃ = Re (ϕ̃c e i θ ), θ is the GO phase, and
.
ωr = −∂t θ is the (real) GO frequency.
In the special case when ωr k·v, Φ depends only on t and x and thus acts as an effective potential.
Such Φ which tends to expel particles from a high-frequency field. In more general settings, though
(for example when a particle interactions with a transverse waves in a strong magnetic field [71]), Φ
can be attractive as well. The concept of dressed particles is also widely known in atomic physics,
where the ponderomotive force −∇Φ is commonly called the dipole force [72].
For such waves, the function J serves as the phase-space density of the wave action, if normalized
properly. One can show (Box 8.1) that it satisfies the so-called wave-kinetic equation (WKE) [41, 69]:
∂J ∂ωr ∂J ∂ωr ∂J
+ · − · = 2ωi J, (17.18)
∂t ∂k ∂x ∂x ∂k
which is a generalization (and correction) of Eq. (16.26).2 Then one has Ref. [67, 69]
∂t d p F + ∇ · d p V F = 0, (17.19a)
2 The WKE can be understood as the Vlasov equation for wave quanta. Alternatively, the Vlasov equation for plasma
particles can be considered as the classical (GO) limit of the WKE derived for particles as quantum waves.
143
LECTURE 17. DRESSED PARTICLES AND PONDEROMOTIVE FORCES 144
d p pF + d k kJ + ∇ · d p pV F + d k kvg J + ∇ d p ΦF = 0,
∂t (17.19b)
d p H0 F + d k ωr J + ∇ · d p H0 V F + d k ωr vg J + ∇ · d p V ΦF = 0,
∂t (17.19c)
. .
where H0 = p2 /(2me ) is the energy of a free OC and v g = ∂k ωr is the group velocity. Hence, equations
of quasilinear theory conserve the particle (OC) number, the total momentum, and the total energy:
d x d p F = const, (17.20a)
d x d p pF + d x d k kJ = const, (17.20b)
d x d p H0 F + d x d k ωr J = const. (17.20c)
The density of the energy–momentum density of the OC distribution is different from that of the
particle distribution precisely by the difference between the wave energy-momentum (Lecture 5) and
the field energy-momentum (which is why electrostatic waves can carry momentum while electrostatic
fields cannot):
In particular, as F diffuses, the OC momentum changes, but so does the wave momentum, and the
total momentum d v me vf0 remains conserved. Also note that in general, the ponderomotive force
on plasma is the sum of all the terms that are quadratic in Ẽ in Eq. (17.19b).
144
Appendices for Part IV
145
APPENDICES FOR PART IV 146
Figure 17.1: Properties of the plasma dispersion function Z (copied from Ref. [73]).
146
Problems for Part IV
Hint: This is doable even without representing the remaining integral through ele-
mentary functions (which is also possible [74] but not required here).
147
PROBLEMS FOR PART IV 148
(a) Assuming k > 0, calculate the response function σk (t) and identify the characteristic time scale
of phase mixing of such plasma. Take the Laplace transform of σk (t) to obtain the spectral
representation of the conductivity, σ(ω, k).
(b) Using the result from part (a), calculate χ(ω, k).
(c) Show that the same result is obtained using Landau’s rule and
ωp2 0
χ(ω, k) = − 2
k
d v v f−0 (v)
ω/k
. (17.26)
L
(e) Derive the general expression for σ(t, x) in terms of f0 (v). Plot the result for the Lorentzian
plasma. Explain the plot qualitatively.
S −iD 0
ω2
= iD S 0 , S ≈ Ωpi2 1, D ≈ − Ωω S S . (17.27)
0 0 P i i
(a) Assume that P is large enough such that E z is negligible. Starting from Maxwell’s equations
.
and assuming the notation Nz = ckz /ω, show that E y satisfies
c d D
2 2 2
+ S − Nz −
2
ω 2 d x2 S − Nz2 E y = 0. (17.28)
(b) Assume S = (1 + x/L)Nz2 , so that x = 0 corresponds to the so-called Alfvén resonance, where
S = Nz2 . Assume L > 0. Argue that Eq. (17.28) can be written as
d 2
x2 − 1
x 2+ E y = 0, (17.29)
dx 2α2
where x has been appropriately rescaled and α is a constant that you are asked to find. Plot
the GO dispersion curves kx (x) ≶ 0 corresponding to Eq. (17.29). Also find and plot the inverse
function, which has the form x(kx ) = x̄(kx ) ± ∆(kx ).
(c) In the GO limit, calculate the x-component of the group velocity at S − Nz2 D (for x > 1)
and S − Nz2 D (for −1 < x < 0) to determine the direction of the action flows along the
dispersion curves. Identify the branches at x → +∞ and at x → 0−. Describe what happens to
a wave launched toward the Alfvén resonance.
148
PROBLEMS FOR PART IV 149
.
(d) Consider g1 (kx ) = e i Θ(kx ) E y (x)e i kx x d x, where Θ(kx ) = x̄(kx ) d kx and x̄ is the same as in
.
problem (b). Take for granted (or prove it yourself) that g1 and g2 = −i sg1 − g10 satisfy
.
i ddk gg1 = si −s −i
g1 .
, s = α2 kx2 . (17.30)
x 2 g2
Assume that g1,2 ∝ e −i θ(kx ) , where ζ = θ0 (kx ) changes slowly with kx compared with θ. Derive
.
the two corresponding GO dispersion branches ζ1,2 (kx ) and explain how they are related to
x(kx ) in problem (b). By considering the “frequency gap” |ζ1 − ζ2 |, estimate the parameter that
determines the mode-conversion efficiency. Is energy deposition at the Alfvén resonance larger
at small α or at large α?
(a) Assume that f0 (v) is bi-Maxwellian with zero average velocity, namely,
!
1 vx2 vy2 vz2
f0 (v) = 3/2 2 exp − 2 − 2 − 2 , (17.32)
π w⊥ wk w⊥ w⊥ wk
2
where w⊥ = 2T⊥ /m and wk2 = 2Tk /m. Show that the dielectric tensor is diagonal. Then
calculate ⊥ explicitly and show that the dispersion relation of transverse waves can be written
as follows:
T⊥
ω 2 − k 2 c2 − ωp2 + ωp2 [1 + ζZ(ζ)] = 0. (17.33)
Tk
[If needed, the latter term can as well be represented in terms of Z 0 (ζ).]
(b) Briefly comment on how to recover the cold-wave dispersion relation (10.20) from Eq. (17.33).
(c) For the hot-plasma limit (ζ 1), show that Eq. (17.33) is approximately linear in ω and derive
the corresponding ω(k) explicitly. (For simplicity, assume k > 0.) Show that, at T⊥ > Tk , there
always exist k for which waves are unstable.
149
PROBLEMS FOR PART IV 150
(d) For the cold-plasma limit (ζ 1), assume that Im ζ > 0 and that Re ω is small (which can be
checked a posteriori ); then the plasma dispersion function can be approximated as3
1 1
Z(ζ) ≈ − − 3 .
ζ 2ζ
Assuming this asymptotic, show that Eq. (17.33) is biquadratic in ω (i.e., quadratic in ω 2 ).
Calculate ω(k) to the first order in the temperature and show that one of the modes is unstable.
Show that T⊥ Tk is required to justify the cold-plasma approximation for this mode.
(e) Compare your analytic calculations with numerical solution of Eq. (17.33).
2 t−t0
ms ωps
s (t, x) = −e dp v d τ eiβ
j (i) i k·x−i ωt Ẽx U cos(φ + Ωs τ )
4π 0
∂f0s
+ Ẽy U sin(φ + Ωs τ ) + Ẽz − V cos(φ + Ωs τ ) , (17.34)
∂pk
k⊥ v⊥
β=− [sin(φ + Ωτ ) − sin φ] + (ω − kk vk )τ. (17.35)
Ω
Hint: For the remaining notation, see the slides from Lecture 13. In particular, re-
. .
member that Ωs = Ωs0 /γ is the relativistic gyrofrequency and Ωs0 = es B0 /(ms c) is the
nonrelativistic gyrofrequency. Also, f0s is the momentum distribution, not the velocity
distribution; the two are connected by the factor m3 , as explained in Sec. 8.1.2.
The derivation can be performed as explained on the slides, but you may also simplify
some of the steps, because (χs )xx is somewhat easier to calculate than the whole tensor.
(b) Assuming that f0s is isotropic, nonrelativistic, Maxwellian, and has zero average velocity, show
that
2 ∞
ωp0 n2 In (λs ) −λs
e An , An = k w . Z(ξn )
X
(χs )xx = (17.37)
ω n=−∞ λs k s
3 Note that this asymptotic is slightly different from the asymptotic of Z on the real axis derived in Sec. 11.2.1. See
150
PROBLEMS FOR PART IV 151
Present the corresponding dispersion relation and identify the corresponding waves and their
polarizations.
(b) By retaining the lowest-order finite temperature corrections, show that the dispersion relation
for Alfvén waves can be written as
pk − p⊥
ω 2 = kk2 VA2 1 − 2 , (17.39)
B0 /(4π)
where pk and p⊥ are the parallel pressure and the perpendicular pressure, respectively. This
dispersion relation predicts the an instability, which is known as the firehose instability. When
does this instability occur? Qualitatively, what is its mechanism?
Hint: See slides for the formulas for An and Bn . Neglect the contribution from
the resonant pole but keep two terms when expanding the rest of the plasma disper-
sion function. Then simplify the result by adopting that ω Ωi . However, when
expanding in ω, assume that the parameter kwk /ω is fixed and of order one.
(a) Using Eqs. (17.38) with T⊥ = Tk , derive the whistler-wave dispersion relation. Unlike in Prob-
lem PIV.7, neglect thermal effects in Re R̄ but account for nonzero Im R̄.
(b) Present the dispersion relation for ωr . Show that in the regime when N 2 1 and ωr |Ωe |,
√
the group velocity of these waves scales as vg ∝ ωr .
(c) Suppose that a whistler wave is excited by an antenna with fixed real frequency. Estimate the
distance that the wave propagates before it damps assuming that damping is weak.
Hint: The idea is the same as in Sec. 10.2.3 except: (i) the dispersion function is
different and (ii) complex is the wavevector rather than the frequency.
151
PROBLEMS FOR PART IV 152
(b) Suppose that the static magnetic field is inhomogeneous, B0 ≈ B̄0 (1 + z/`0 ). Here, z is the
coordinate along the field (the wave propagates from z = −∞ to z = +∞), `0 is some large
enough characteristic length (you may assume that `0 > 0 for simplicity, but this is not essential),
and the constant B̄0 is the field strength at which the wave is in cyclotron resonance with some
minority ions s. Assume that the wave frequency ω is fixed and given. Also, assume that the
wave amplitude |Ẽ± | is given at z = 0 (and depends on z slowly enough such that its value
outside the resonance region is unimportant). Calculate the total wave power d z PCD,s that
is dissipated through cyclotron damping on the minority ions per unit cross section of the wave
beam using two different methods (make sure that the results are consistent):
(i) using Eq. (15.26);
(ii) using Eq. (15.10) with the cold-plasma dielectric tensor (6.13).
(c) Propose an analytic estimate for the characteristic length of the absorption region aabs . Assum-
ing parameters typical for ion cyclotron heating in tokamaks, estimate aabs numerically.
152
Appendices
153
Appendix A
Abbreviations
Abbreviation Meaning
GO geometrical optics
CCS collisionless cold static
DK Dnestrovskii–Kostomarov
EBW electron Bernstein wave(s)
EPW electron plasma wave(s), a.k.a. (electron) Langmuir waves
IAW ion acoustic wave(s), a.k.a. ion sound waves
IBW ion Bernstein wave(s)
IPW ion plasma wave(s)
MHD magnetohydrodynamic(s)
OC oscillation center(s)
ODE ordinary differential equation
PDE partial differential equation
TE, TM transverse-electric, transverse-magnetic
WKB Wentzel–Kramers–Brillouin
WKE wave-kinetic equation
WWT Wigner–Weyl transform
154
Appendix B
Notation
This appendix summarizes the notations that are of general importance or often used in the main text.
∂ partial derivative
∂a ∂/∂xa partial derivative with respect to a spatial coordinate
∂
α ∂/∂xα partial derivative with respect to a spacetime coordinate
principal-value integral
1 unity
1 unit matrix
1̂ unit operator
1̂ unit matrix operator
155
APPENDIX B. NOTATION 156
B, Ba magnetic field
B 0 , B0 background magnetic field
B , Ba magnetic field envelope
CS Eq. (7.33) ion sound speed
D Eq. (6.14b) element of of CCS magnetized plasma
D, Dab symbol of a dispersion operator
D̂, D̂ab dispersion operator
Dϕ electrostatic dispersion operator
D̂ E Eq. (1.23) electromagnetic dispersion operator
E, Ea electric field
E , Ea electric field envelope
H
~ω photon energy or photon Hamiltonian
I dx I wave action
I Sec. 5.1 wave action density
Im imaginary part
J , Ja vg I action flux density
L Eq. (6.14d) element of of CCS magnetized plasma
Lc characteristics spatial scale of an envelope and (or) of a medium
L Sec. 10.1.2 Landau contour
N ck/ω refractive index
N dimension of a wave field (N = 3 for the electric field)
P Eq. (6.14c) element of of CCS magnetized plasma, also pressure
P χ̂E electric polarization
P kI wave momentum density
Pabs absorbed power
R Eq. (6.14d) element of of CCS magnetized plasma
156
APPENDIX B. NOTATION 157
Q, Q Stern–Gerlach term
Re real part
S Eq. (6.14a) element of of CCS magnetized plasma
S Eq. (5.17) Poynting vector
Tc characteristics temporal scale of an envelope and (or) of a medium
Ts temperature of species s
U ωI wave energy density
Vα ∂D/∂kα k-derivative of the symbol of a dispersion operator
W , W −1 Sec. 3.3 direct and inverse Wigner–Weyl transform
Zs es /e charge state of species s
157
Appendix C
Conventions
This appendix summarize the mathematical conventions assumed in the main text.
z = zr + i zi , z ∗ = zr − i zi . (C.2)
Similarly, for any given matrix or operator M , its Hermitian part MH and anti-Hermitian part MA
are defined as
. 1 . 1
MH = (M + M † ) = MH† , MA = (M − M † ) = MA† . (C.3)
2 2i
Accordingly,
M = MH + i MA , M † = MH − i MA . (C.4)
C.2 Geometry
Euclidean or pseudo-Euclidean coordinates are assumed. The bold font is used for spatial coordinates
(x = {x1 , x2 , ...xn }, with any n ≥ 0), vectors (covectors), and matrices. The upper and lower indices
of these objects are marked with Latin indices and are interchangeable; for example, ka = k a , with
a = 1, 2, . . . n. For any two n-dimensional column vectors X and Y , one has
.
X · Y = X † Y = Xa∗ Y a , (C.5)
158
APPENDIX C. CONVENTIONS 159
The spacetime coordinates and spacetime vectors (covectors) are denoted with the sans-serif font
in the invariant representation; but this font is used also for other objects. For example, x = {x0 , x},
.
and n = dim x = n + 1. The corresponding components are denoted with a regular font and Greek
indices (for example, xα ), which span from 0 to n. For the spacetime, Minkowski metric is assumed
in the form gαβ = g αβ = diag {−1, 1, 1, . . .}. For the objects in the Minkowski space, conventions
are similar to those above, except the indices are raised and lowered by the Minkowski metric. For
example, kα = gαβ k β , or equivalently, k α = g αβ kα , so in particular, k 0 = −k0 and k a = ka for a > 0.
159
Bibliography
160
[20] D. E. Ruiz, Geometric theory of waves and its applications to plasma physics, Ph.D. Thesis,
Princeton University (2017), arXiv:1708.05423.
[21] Yu. A. Kravtsov, L. A. Ostrovsky, and N. S. Stepanov, Geometrical optics of inhomogeneous and
nonstationary dispersive media, Proc. IEEE 62, 1492 (1974).
[22] M. Bornatici and Yu. A. Kravtsov, Comparative analysis of two formulations of geometrical
optics. The effective dielectric tensor , Plasma Phys. Control. Fusion 42, 255 (2000).
[23] M. Bornatici and O. Maj, Geometrical optics response tensors and the transport of the wave
energy density, Plasma Phys. Control. Fusion 45, 1511 (2003).
[24] G. B. Whitham, Linear and Nonlinear Waves (Wiley, New York, 1974).
[25] K. Y. Bliokh, A. Niv, V. Kleiner, and E. Hasman, Geometrodynamics of spinning light, Nature
Phot. 2, 748 (2008).
[26] R. G. Littlejohn and W. G. Flynn, Geometric phases in the asymptotic theory of coupled wave
equations, Phys. Rev. A 44, 5239 (1991).
[29] L. D. Landau and E. M. Lifshitz, The Classical Theory of Fields (Pergamon Press, New York,
1971).
[30] W. D. Hayes, Group velocity and nonlinear dispersive wave propagation, Proc. R. Soc. Lond. A
332, 199 (1973).
[31] D. E. Ruiz and I. Y. Dodin, On the correspondence between quantum and classical variational
principles, Phys. Lett. A 379, 2623 (2015).
[32] L. Landau, Zur Theorie der Energieubertragung. II , Phys. Z. Sowjetunion 2, 46 (1932).
[33] C. Zener, Non-adiabatic crossing of energy levels, Proc. R. Soc. London A 137, 696 (1932).
[34] E. R. Tracy and A. N. Kaufman, Metaplectic formulation of linear mode conversion, Phys. Rev.
E 48, 2196 (1993).
[35] T. E. Stringer, Low-frequency waves in an unbounded plasma, J. Nucl. Energy C 5, 89 (1963).
[36] L. D. Landau and E. M. Lifshitz, Statistical physics (Pergamon Press, New York, 1980), pt. 1.
[37] C. R. Sovinec, Note on kinetic Alfvén waves, UW-CPTC 09-2, https://cptc.wiscweb.wisc.edu/
wp-content/uploads/sites/327/2017/08/UW-CPTC 09-2 rev.pdf.
[38] E. W. Weisstein, Abel Transform, from MathWorld – A Wolfram Web Resource, https://
mathworld.wolfram.com/AbelTransform.html.
[39] B. Eliasson and K. Papadopoulos, HF wave propagation and induced ionospheric turbulence in
the magnetic equatorial region, J. Geophys. Res.-Space 121, 2727 (2016).
[40] J. E. Moyal, Quantum mechanics as a statistical theory, Proc. Cambridge Philosoph. Soc. 45, 99
(1949).
[41] S. W. McDonald and A. N. Kaufman, Weyl representation for electromagnetic waves: The wave
kinetic equation, Phys. Rev. A 32, 1708 (1985).
[42] http://www.princeton.edu/∼idodin/papers/APS-DPP-CPP.pdf.
[43] R. W. Gould, T. M. O’Neil, and J. H. Malmberg, Plasma wave echo, Phys. Rev. Lett. 19, 219
(1967).
[44] N. G. Van Kampen, On the theory of stationary waves in plasmas, Physica 21, 949 (1955).
[45] K. M. Case, Plasma oscillations, Ann. Phys. (N.Y.) 7, 349 (1959).
[46] E. C. Taylor, Landau solution of the plasma oscillation problem, Phys. Fluids 8, 2250 (1965).
[47] L. Landau, On the vibration of the electronic plasma, J. Phys. USSR 10, 25 (1946).
[48] J. H. Malmberg and C. B. Wharton, Collisionless damping of electrostatic plasma waves, Phys.
Rev. Lett. 13, 184 (1964).
[49] R. K. Mazitov, Damping of plasma waves, Zh. Priklad. Mekh. Tekh. Fiz. 1, 27 (1965).
[50] T. O’Neil, Collisionless damping of nonlinear plasma oscillations, Phys. Fluids 8, 2255 (1965).
[51] P. F. Schmit, I. Y. Dodin, and N. J. Fisch, Evolution of nonlinear waves in compressing plasma,
Phys. Plasmas 18, 042103 (2011).
[52] I. B. Bernstein, J. M. Greene, and M. D. Kruskal, Exact nonlinear plasma oscillations, Phys.
Rev. 108, 546 (1957).
[53] I. Y. Dodin, On variational methods in the physics of plasma waves, Fusion Sci. Tech. 65, 54
(2014).
[54] C. Liu and I. Y. Dodin, Nonlinear frequency shift of electrostatic waves in general collisionless
plasma: unifying theory of fluid and kinetic nonlinearities, Phys. Plasmas 22, 082117 (2015).
[55] W. L. Kruer, J. M. Dawson, and R. N. Sudan, Trapped-particle instability, Phys. Rev. Lett. 23,
838 (1969).
[56] I. Y. Dodin, P. F. Schmit, J. Rocks, and N. J. Fisch, Negative-mass instability in nonlinear plasma
waves, Phys. Rev. Lett. 110, 215006 (2013).
[60] P. A. Robinson, Dispersion of electron Bernstein waves including weakly relativistic and electro-
magnetic effects. Part 1. Ordinary modes, J. Plasma Phys. 37, 435 (1987).
[61] P. A. Robinson, Dispersion of electron Bernstein waves including weakly relativistic and electro-
magnetic effects. Part 2. Extraordinary modes, J. Plasma Phys. 37, 449 (1987).
[62] I. Y. Dodin and A. V. Arefiev, Parametric decay of plasma waves near the upper-hybrid resonance,
Phys. Plasmas 24, 032119 (2017).
[63] R. I. Pinsker, Whistlers, helicons, and lower hybrid waves: The physics of radio frequency wave
propagation and absorption for current drive via Landau damping, Phys. Plasmas 22, 090901
(2015).
[64] M. Porkolab, On the physics of magnetosonic wave damping on electrons, AIP Conf. Proc. 244,
197 (1992).
[65] A. A. Vedenov, E. P. Velikhov, and R. Z. Sagdeev, Nonlinear oscillations of rarified plasma, Nucl.
Fusion 1, 82 (1961).
[66] W. E. Drummond and D. Pines, Non-linear stability of plasma oscillations, Nucl. Fusion Suppl.
3, 1049 (1962).
[67] R. L. Dewar, Oscillation center quasilinear theory, Phys. Fluids 16, 1102 (1973).
[68] S. W. McDonald, C. Grebogi, and A. N. Kaufman, Locally coupled evolution of wave and particle
distribution in general magnetoplasma geometry, Phys. Lett. A 111, 19 (1985).