(Dwight R Nicholson) Introduction To Plasma Theory PDF
(Dwight R Nicholson) Introduction To Plasma Theory PDF
|0,| and wp > w;. However, for w; and |Q,), there are two possibilities: w, >
|0,|, and w, < [,|.
EXERCISE Show that w, = |Q,| when w, = 24/01.
The dispersion diagrams are different in the two cases. These are found in Ref. (3),
p. 195, The dispersion relation w = a{k) is shown in Fig. 7.17 for the case
w, > |0,). Note that the R-wave has two “pass bands,” 0 < w < {0,|, and
w > wp, Separated by a “stop band.” The L-wave exists only for w > w,. Both
high frequency branches asymptote to w ~ ke at high frequencies. The locations
of the pass and stop bands can be seen more clearly by drawing n? = k?c*/w* vs.
w, as shown in Fig. 7.18 (see Ref. (3), p. 194), The low frequency branch of the
R-wave is often called the electron-cyclotron wave. Once again, the “stop bands”
occur where n? < 0 and the “pass bands” occur when n? > 0. (Do we trust this
theory for low frequencies w < ,?)
For the low density plasma, w, < |Q,|, the character of the dispersion relation
«@ = w(k) changes, as shown in Fig. 7.19 (see Ref. (3), p. 195). The low frequency
branch of the R-wave is again called the electron-cyclotron wave.
k
Fig. 7.17 Dispersion diagram for parallel electromagnetic waves for the case w, > (0,|.160 Fluid Equations
|
!
1
1
t
|
(
Fig. 7.18 Dispersion diagram for parallel electromagnetic waves for the case w, > |Q,\.
From Fig. 7.19 we can see that the electron-cyclotron wave has a portion where
V, = dw/dk increases as w increases. This is called the whistler wave, because the
high frequency components of a wave packet travel faster than its low frequency
components, An observer some distance away from a source (a lightning stroke,
for example) will then hear a whistle starting at high frequencies and descending to
lower frequencies.
In both of our dispersion diagrams, the R-wave at very high frequencies is seen
to have a higher phase speed than the L-wave. Thus, ifa plane wave is incident on
R (electron cyclotron wave}
Fig. 7.19 Dispersion diagram for parallel electromagnetic waves for the case w, < (0.1Alfven Waves 161
a plasma along Bp, its two normal mode components, R and L, travel at different
speeds, and the plane of polarization of the plane wave rotates as it travels. This is
known as Faraday rotation, and is useful in measuring plasma densities in labora-
tory plasma and in interstellar space
This completes our discussion of high frequency electromagnetic waves (ignor-
ing ion motion) traveling in a magnetized plasma. We have discussed only waves
traveling across By (O-mode and X-mode) and along 8, (R-wave and L-wave). Of
course, waves can travel at any angle to By. When they do, there will be two modes
for any angle of propagation, and their properties will be some combination of the
properties of the O, X, R, and L waves.
7.11 ALFVEN WAVES
Up until this point, we have considered electromagnetic waves ignoring ion mo-
tion, and ion waves that were electrostatic and thus ignored electromagnetic ef-
fects. Let us next combine ion motion with electromagnetic effects, we shal! find
parallel Alfvén waves (k || 8)) and perpendicular magnetosonic waves (k 1 8).
First we look for low frequency waves traveling along By. For simplicity we take
acold plasma, T; = T, = 0. We can also ignore electron inertia (m, — 0). Just as
in the case of R-waves and L-waves, we look for waves with k = k2, By = Bo2,
and V,, V,, E,, By all in the x-y plane (Fig. 7.20). Unlike the R-, L-wave case, we do
not look for a rotating E,, B,; rather, we take E, = E,£ and B, = By. As we
shall see, this form for E, and B, is not entirely self-consistent. The relevant fluid
equations are then (linearizing and Fourier transforming immediately):
Fig. 7.20 Vector orientations in an Alfvén wave.162 Fluid Equations
kx Ee, = 28, (7.202)
ik XB, = az nye(V, — V,) — “2B, (7.203)
0 = — em E, — = Vv, x By (7.204)
s Ny
iam, = enoE, + “2 V, x By (7.205)
Considering the motions of individual electrons, we are keeping the electron E, *
B, drift in the 9-direction, while ignoring the polarization drift in the £-direction.
We recall from Chapter 2 that the polarization drift speed is proportional to mass.
Thus, for the ions, we have an E, By drift in the f-direction, which approxi-
mately equals the electron drift and prevents any current in the J-direction. We
also have for the ions a component of velocity in the ¥-direction (a polarization
drift) that provides the current 1)e(V, — V,) in (7.203) to drive the magnetic field
B,. The approximation in this derivation is to ignore that portion of V,, due to the
Vi, & % By force.
With this introduction, we write the relevant components of (7.202) to (7.205) as
oe :
fhe g, = “tm y, - es, (7.206)
o ¢ c
—iwmVy, = eB, + £ VyBo (7.207)
and
~tomyY, = — = v,B (7.208)
Solving (7.208) for V;,, we find
y,=- ify, (7.209)
@
We insert (7.209) in (7.207) to obtain
e/m;
Me =(=o7 =the) E, (7.210)
Combining (7.210) and (7.206) then yields the dispersion relation
(7.211)
Enforcing the assumption @ << @,, implicit in the above discussion, we ignore
a? << 0} in the second denominator, finding
Ret ket
2 = —— = ——_ —
w= TF ai? ~ 1+ 4np,c7 BF (7.212)
where p, nom, is the ion mass density. If we define the Alfvén speed V, =
(Be/4109)%, (7-212) isAllven Waves 163
Kec?
2 BrecsCecoeeser
# = TT ewe (7.213)
Multiplying top and bottom by V,,2/c? we finally obtain
RV.
ae
oF Tt 7A Cae
as the dispersion relation for Alfvén waves. Note that for V, << c¢, this isw =
kV ,; therefore we have an acoustic dispersion relation. Recall that acoustic waves
in air have an acoustic speed (P/p,,) where P is pressure; here we have a speed
V, = (B2/4mp,,.)", which is of the same form if we relate By?/4m to a magnetic
pressure.
The physical interpretation of Alfvén waves is very interesting. We have seen
that electrons andions are E, X By drifting together in the 9-direction, with speed
—(E,/B,)c. Thus, both plasma fluids move together in the J-direction. Now what
is happening to magnetic field lines? They are being distorted by the addition of
B, = B,f to the background magnetic field By = Bo2, as shown in Fig. 7.21. The
position function of a magnetic field line can be defined as
Y4(z.1) =f 2¢g2 de’ (7.213)
Then the 9 velocity V,, of a magnetic field line is the time derivative of Y,, or {with
By ~ exp (—iwt + ikz)]
= B,
=-j a a
¥,=-iwf Ba (7.216)
Now from (7.202) and Fig, 7.20 we have
a= he, (7.217)
or
Vy = ~(Ex/By)c (7.218)
which is precisely the $-velocity of fluid flow. Thus, in the J-direction, we say that
the particles are frozen to the field lines. This concept will prove useful later
Fig. 7.21 Total magnetic field in an Alfvén wave.164 Fluid Equations
(Chapter 8) for other low frequency plasma motions. Note that in the #-direction,
we can take the fluid speed and the field line speed both to be zero, satisfying this
concept. However, in the £-direction, we have seen that V,, ~ 0 while V,, # 0.
Thus, we cannot have both kinds of particles frozen to the field lines in the
&-direction.
One may recall that the wave equation of a stretched string is w = key with
cr = VT/p,, where T is the tension on the string and p,, is the mass per unit
length. If we identify B?/4m as a tension per unit area and p, as a mass per unit
volume, then the Alfvén wave dispersion relation w = kV, can be thought of as
representing the wave that propagates when a field line, loaded with plasma, is
plucked in the transverse direction,
7.12 FAST MAGNETOSONIC WAVE
The Alfvén wave of the previous section is a low frequency parallel electromagnetic
wave, traveling along 8. Let us now look for a low frequency perpendicular
electromagnetic wave, traveling across By; this is the fas! magnetosonic wave.
For simplicity, consider a cold plasma, T, = T; = 0, and ignore electron iner-
tia, m, > 0. We look for a wave withk- B, = 0,k+E, = 0, andE,-B, = 0,as
in Fig. 7.22.
EXERCISE Why don’t we look for low frequency waves with E, along By?
Have we ever looked for a wave with E, along B,? What did we find?
We choosek = k9,E, = E,%,andB, = B,2; thus B, is along B,, and the relevant
fluid equations, linearized and Fourier transformed, are
ik x E, = fe B, (7.219)
#
Fig. 7.22 Vector orientations in a fast magnetosonic wave.Fast Magnetosonic Wave 165
x B, = “7 ney, - v,) - 2 &, (7.220)
0 =— enE, — _ Vv, x B, (7.221)
; ey
siwmnV, = en, + 2 v, x By (7.222)
From (7.221) we see that the electrons will have only an E, X By drift in the
&-direction, while the ions, for very small w, will have approximately the same
E, X Bo drift in the £-direction. The ions, however, will have an extra component
of velocity in the £-direction, along £,, because of the polarization drift, which
produces a current in the £-direction that produces the perturbed magnetic field in
the 2-direction. The relevant components of (7.219) to (7.222) are then
~IkE, = e B, (7.223)
ikB, = Arve we (7.224)
iwmV, = eE, + £ V,Bo (7.225)
iamViy = V,,Bo (7.226)
These equations are identical to (7.206) to (7.209) for the Alfvén waves, and we
can immediately write the dispersion relation (7.214) for small frequencies; this is
RV2
TF He (7.227
aw’
for the fast magnetosonic wave traveling across By.
This completes our discussion of linear wave equations in infinite uniform
plasma. We have often looked at parallel and perpendicular waves; in practice,
waves can propagate at any angle to the magnetic field. Waves propagating at an
arbitrary angle usually have some combination of the properties of the corre-
sponding parallel wave and the corresponding perpendicular wave. Because of the
complexity of these waves, we shall not derive all of their properties here. How-
ever, a useful qualitative device exists for thinking about these waves. This is called
the CMA diagram, after its inventors Clemmow, Mullaly, and Allis (4, 5]. The
diagram is valid only for cold plasmas, T; = T, = 0. It shows all of the waves
that can propagate at a given angle to the magnetic field for any combination of
density and magnetic field intensity. This useful diagram is discussed in Refs. [6]
and (7).
In the next section, we turn our attention from linear waves characterized by a
real frequency and a real wave number, to linear instabilities characterized by a
complex frequency and a real wave number,166 Fiuld Equations
7.13 TWO-STREAM INSTABILITY
Previous sections of this chapter have treated examples of linear waves that arise
from the fluid theory of plasma. These waves are characterized by a real frequency
and a real wave number, and would all be excited if a magnetized Maxwellian
plasma were perturbed. When a plasma does not consist of Maxwellian electrons
and Maxwellian ions, some of the waves (normal modes) of the system can become
unstable. This subject is treated within the Vlasov theory in Section 6.9, Within the
fluid theory, unstable normal modes can arise whenever the zero order electron
and ion velocities are different, or whenever one species consists of two or more
components each with different zero order velocities. Such instabilities are called
streaming instabilities.
As an example of a streaming instability, consider a plasma in which the ions
are stationary, while the electrons are traveling with speed Vy, The linearized fluid
equations are then
OMe + Mo Vey + Vo AM = 0 (7.228)
met Ve, + meMeVo 8,Ver = ~ engE (7.229)
8,n, + 9 AV = 0 (7.230)
myfty 0, Vin = eMoE (7.231)
and
a,£ = 4re(ny ~ na) (7.232)
where we have assumed one-dimensional motions and T, = T, = 0; and the
zeroth order speed V, contributes an extra term in (7.228) and in (7.229). Because
we have no reason to suspect that the oscillations found here will be low frequen-
cy, we keep Poisson’s equation (7.232) and we do not assume quasineutrality. In
fact, if we allow m, — © in (7.231), we would simply obtain the drifting cold
plasma waves discussed in Problem 7.4; these are high frequency waves that
become Langmuir waves in the limit that the drift speed V, ~ 0. Here, we keep m,
large but finite and show that the drifting cold plasma waves are unstable; that is,
when the frequency w is obtained from (7.228) to (7.232), one finds Im(w) > 0;
thus exp (—iwt) ~ exp [Im(w)s], which grows exponentially with time. Since no
instability is found in Problem 7.4, it must be the case that Im(w) — Qasim, — 2
Fourier transforming (7.228) to (7.232), Eq. (7.229) yields
(“iw + ikV,)Vq = — cE/m, (7.233)
which when inserted in (7.228) yields
oe _thngeE/me
(le + HK ,)m, — A = 0 (7.234)
while (7.231) in (7.230) yields
~tknyeE/
~ien, = (7.235)
Then using (7.234) and (7.235) in Poisson's equation (7.232), we find the disper-
sion relationTwo-Stream Instability 167
iknge )
eae (7.236)
or
ek,w) (7.237)
where we have identified the dielectric function «(k,w) (see Section 7.4). In the
limit m, — %, w, — 0, we find
w= k¥, + w, (7.238)
in agreement with Problem 7.4.
With m; finite, Eq. (7.237) is a quartic equation in w, with four roots. Since
(7.237) is a real equation, the complex conjuate of any root is also a root. (Why?)
Thys, if we find any complex roots, either that root or its complex conjugate wil!
have Im(w) > 0, and there will be an instability.
Let us use enlightened guessing to solve (7.237), Since the ions are important, we
look for a wave such that the frequency is low in the laboratory frame [e.g.,
kV, ~ w, and the lower sign in (7.238)]. However, low frequency means only
|| << w,; a vigorous instability could well lead to |w| >> «,;. Let us then look
for a solution (possibly complex) to (7.237) that satisfies a; << |wl << a.
Then, because the second term in (7.237) is much less than unity, in order to cancel
the first term we must have the third term close to unity; this leads us to look at
wave numbers & such that kV = w,. Then (7.237) yields
(7.239)
or
= + wPw, (7.240)
or
a "
) { ) (7.241)
which represents instability since one of the three values of (—1)'” is (1/2) +
i(3/2). In the frame moving with the electrons, the Doppler shifted frequency is
w! = w — kV; since KV) = w, and |w| << w, this is roughly w ~ — w,, so
that the electrons see an oscillation at nearly their natural frequency of oscillation.168 Fluld Equations
There is another useful way to determine that (7.237) yields instability. From
(7.237) we define
2
Voy
wo?
we
F(k,a) (7.242)
(w
We can plot this function versus real frequency w at fixed wave number k, as
sketched in Fig. 7.23. From (7.242) and the illustration we see that when the line at
unity intersects the graph of F(k,w) at four different points, there are four real
roots and no instability for the chosen value of k. However, suppose the central
minimum of F(x,w) occurs at a value greater than unity; then there are only two
real roots, as shown in Fig. 7.24. To determine when this happens, we determine
when
Frin(ky@o) > 1 (7.243)
where Fnin is determined by 0F/dw = 0 from (7.242). We find
pce ()"er. (7.248)
and
(7.245)
RV?
which satisfies (7.243) and predicts instability whenever
|kVol = a, (7.246)
Thus, there is a broad range —w,/V>,< k < w,/Vo of unstable wave numbers.
Two-stream instabilities are very common in plasma physics. They happen
whenever one fairly cold plasma component has a relative velocity with respect to
another plasma component. These components need not be of different species; a
cold electron beam impinging on an existing electron-ion plasma will produce its
own instability. These instabilities are nature’s way of saying that Maxwellians are
desirable, and any configuration that is too far from Maxwellian will not last
forever, even in the absence of collisions.
The linear theories of streaming instabilities for both cold components (fluid
theory) and warm components (Vlasov theory) are very well understood. The
F(k,w)
2
0 #,
Fig. 7.23 Graphical solution of (7.237), for wave numbers k that yield four real roots.Orift Waves 169
Orin Vy
min kV,
Fig. 7.24 Graphical solution of (7.237), for wave numbers k that yield only two real roots.
nonlinear saturation of these instabilities, involving such concepts as particle orbit
modification, nonlinear wave-wave interactions, and strong turbulence, are not so
well understood, and are the subject of considerable current research.
Up to this point in our study of the fluid theory, the waves and instabilities have
propagated in a spatially homogeneous plasma. In the next section, we consider
waves that propagate in a spatially inhomogeneous plasma.
7.14 DRIFT WAVES
Spatial inhomogeneities can give rise to their own wave motions. Consider an
electrostatic wave, with frequency high enough that ions are unperturbed, but low
enough that electrons perform an E X By drift in the wave field (Fig. 7.25). The
wave number is predominantly in the f-direction, but has a small 2 component to
allow electrons to flow freely along the field lines. With E, = E,f + E.2, the
E,f X Bo2 drift is in the £-direction, causing a charge separation. The continuity
equation is
an + V-(aVy=0 (7.247)
or
sian, + anViy = (7.248)
Fig. 7.25 Vector orientations for an electrostatic drift wave.170 Fluld Equations
where we assume V,, is not a function of x, we ignore the k,V,, term because k, is
small, and we ignore the k,V,, term because V,, is small, being mostly a result ofa
polarization drift. Since V;, results primarily from the E X B drift, we have
— Be
m= 5 (7.249)
since E, ~ E;; thus (7.248) yields
1 am Bi
"Ta ox B * (7.250)
Now the force equation in the 2-direction, ignoring —iwm,noV,, because of the
smailness of w, is
0 — T.ik,my — em,
fw — Dik, — enE(k/k,) (7.251)
or
_ deny
n= Tk (7,252)
or equating (7.250) and (7.252) and eliminating Ey,
2
T.k, am ¢ ve
=~ "ne ox By IO,IE, * (7.253)
where the density scale length
L =-(L Sy >0 (7.254)
: ny de :
Defining the electron diamagnetic drift speed (see Chapter 2)
< (7.255)
we can write (7.253) in the form
(7.256)
which is the dispersion relation for electrostatic drift waves.
There is a whole zoo of drift waves, matching in diversity all of the waves in
homogeneous magnetized plasma. Drift waves are very important in magnetic
confinement devices for controlled fusion such as the tokomak and mirror ma-
chine, and in the study of planetary ionospheres and magnetospheres. They are
discussed in greater detail in Refs. (3], [6], and [8] to [19].
This brings us to the end of our study of /inear fluid waves in magnetized and
unmagnetized, homogeneous and inhomogeneous plasma. In the next two sec-
tions, we introduce the important subject of nonlinear fluid waves by adding one
nonlinear term to two of the most important waves in plasma physics: ion-acoustic
waves and Langmuir waves.Nonlinear lon-Acoustic Waves 71
7.15 NONLINEAR 1ON-ACOUSTIC WAVES—
KORTEWEG-DeVRIES EQUATION
Up to this point in the fluid theory we have considered only linear waves. We must
always remember that the theory of linear waves restricts us to very small ampli-
tudes. A wave with a finite amplitude will be susceptible to nonlinear effects,
which show up mathematically as products of first order terms. This section and
the next section are intended to introduce the concept of nonlinear wave equations
and their corresponding solutions, which often take the form of solitons and shock
waves.
Here we consider an example of one nonlinear wave equation, the Korteweg-
deVries equation [20]:
au + va + adfv = 0 (7.257)
This equation is obtained by adding one nonlinear term in the derivation of the
ion-acoustic wave equation.
Although it is possible to give a rigorous derivation of (7.257), we give here only
a heuristic derivation that indicates how one might arrive at (7.257). The origin of
the terms in (7.257) is fairly easy to see. The first two terms might arise from the
total time derivative of the ion fluid velocity. The third term can be seen in the
ion-acoustic dispersion relation (7.104), which upon taking 7, = 0, y, = 1, is
Ree
= TSE (7.258)
The square root of (7.258) is, for small kA,,
ke, ae RAZ
o= TE BERAE = ke, (1 oe: ) (7.259)
If we now multiply (7.259) on the right by the ion fluid velocity v, and identify —iw
with 0, and ik with 4,, we obtain
av au 2 av
Tag cei cogil ua tl age ‘eae
Ina frame x’ = x — c,f moving with the velocity c,, and defining a = c,A,?/2,
we obtain
dv + adv = 0 (7.261)
which are the linear terms in (7.257). The nonlinear term is obtained by replacing
the partial time derivative 8, with the convective time derivative 0, + vd,
We begin our heuristic derivation with the five fluid equations. Taking 7; —- 0
so that we can neglect ion pressure in the ion force equation, and taking m, > 0
so that we can neglect electron inertia in the electron force equation, we find
an, + a(n.) = 0 (7.262)
0 = -— Tan, — ent (7.263)
an, + anV,) = 0 (7.264)
mn, AV, + mnV, OV; = enE (7.265)
and .
a,£ = ame(n, — 1) (7.266)172 Fluid Equations
where we have chosen y, = 1. We next linearize (7.262) to (7.266) everywhere
except one place: we keep one nonlinear term, the m;n,V; 4,V, term on the left side
of (7.265). We have then
ON + 9 4,V. = 0 (7.261)
0 = ~ T, dng — engE (7.268)
any + My OV, = 0 (7.269)
minty 8,V, + mnoV; 2,V; = enE (7.270)
a,£ = 4me(n, — ma) (7.271)
EXERCISE Can you find seven other nonlinear terms neglected in going from
(7.262)-{7.266) to (7.267)-(7.271)?
A more rigorous derivation would show us the regime of validity implied by our
neglect of seven other nonlinear terms while retaining only one nonlinear term. It
turns out that this regime is reasonably large.
We next assume a plane wave solution, everywhere except in (7.270). [What
would happen if we tried to assume a plane wave solution ~ exp (—iwt + ikx)in
(7.270)?} We also take v = V, * V;, which means that (7.267) and (7.269) have
the same information; we retain the difference between n,, and m,, so that (7.271)
can be used. Solving (7.268) for n,,, we find
=. OME
ta =~ GE (7.272)
which inserted in Poisson’s equation (7.271) yields
el aren,
B= = (wim kT) ae)
tty is from (7.269)
mm = be) 4 (7.274)
Both (7.274) in (7.273) and the result in (7.270) yield
ie? ayaa
ay + vay = — = a+ eaZyy (7.275)
Here, we are still treating the right side as linear; therefore w and k have their
meanings as differential operators, while the left side is nonlinear. It proves conve-
nient to eliminate w on the right side, we do this by using the linear ion-acoustic
dispersion relation (7.258), which is obtained from (7.275) by ignoring the nonlin-
ear term and replacing the left side with —iwv. Solving for w and substituting in
the right side of (7.275), we have
au + vd,v = — ike + AQ e (7.276)
For small kA,, we can expand the right side of (7.276) to obtain
av + vd = — kel — FRA (7.277)Nonlinear lon-Acoustic Waves 173
Reinterpreting ik as 3,, this becomes
av + (c, + v) dv + a Av (7.278)
where a = A,’c,/2. In the frame z = x — ¢,f, this is the Korteweg-de Vries equa-
tion (7.257).
EXERCISE Show the above relationship.
Recall that v(x,1) represents fluid velocity in the laboratory frame; this identifi-
cation of u(x,t) remains true even if we transform to a moving frame. What
physics do the various terms in (7.278) represent? The first two terms by them-
selves,
av + c,av =0 (7.279)
merely represent our old ion-acoustic waves in the limit KA, — 0. The solution of
(7.279) is simply a dispersionless wave, » = ke,, with phase velocity V,, = w/k =
¢,, and group velocity dw/dk = c,a constant independent of k. Suppose we add
the nonlinear term to obtain
Ov t (cp + vy av=0 (7.280)
The effect of the nonlinear term is as follows. Consider an initial waveform as
shown in Fig. 7.26. As the wave moves, the part with larger v moves faster, so that
it overtakes the part with smaller v. Eventually, at¢ = 1, there is an infinite slope,
and at? = ¢;, the wave has broken. Now suppose we had included the dispersive
term in (7.278); the term a 4,°v is called dispersive because it contributes a term k?
to the linear dispersion relation w = kc, — ak?; then V, = dw/dk = c, —
3ak?, which depends on k, making this a dispersive wave. We know the effect of
dispersion on a wave; it makes a wave packet spread out as it travels. This is just
opposite to the steepening observed in the figure. Consider the time between
t = 1, and t = ft). Here, the slope is becoming very large. A large slope corre-
sponds to a large x-derivative, which makes the a @,*v term in (7.278) become
large. Since we know that the effect of this a 0,>v term is to spread out the wave,
we might expect that there could be a balance between the nonlinear steepening
and the linear dispersion. Indeed this is the case. One can obtain nonlinear wave
packets, known as solitons, which travel without change of shape (Fig. 7.27). The
physical basis for these solitons involves a balance between dispersion and non-
linearity.
Fig. 7.26 Effect of the nonlinear term in (7.282).174 Fluid Equations
°
Fig. 7.27. Sketch of a soliton solution.
Let us proceed to find a soliton solution to the Korteweg-deVries equation
(7.278). We look for stationary solutions in a moving frame,
x =x - ut (7.281)
rst (7.282)
so that
8, = (8x'/8x)a,, + (8t'/8x)8, = By (7.283)
and
8, = (8x'/8N)ay + (81'/81)8y = Oy — Wd (7.284)
Since stationary implies 4, = 0, the Korteweg-deVries equation (7.278) becomes
(— uv tc, + v) du + a a,ov =0 (7.285)
Remember that o(x’,1’) is stilt that function of space and time which represents the
fluid velocity in the lab frame. Equation (7.285) can be integrated once immediate-
ly, to give
2
(c, ~ w)v + + av =0 (7.286)
where( ) = 6, ) and we have taken the integration constant to vanish. Equa-
tion (7.286) is in the form
av’ = (w ~ eu — (7.287)
which has the same mathematical form as Newton’s law of motion,
mé = F(x) = — 3,V(x) (7.288)
where V(x) is the potential energy. Thus, (7.287) has the form
- 3
av" = — a, ie = w) $+e] (7.289)
Equation (7.289) has the same mathematical form as a force equation for a parti-
cle of mass a moving under the influence of a potential field given by the quantity
in brackets. We call the quantity in brackets the pseudopotential,
yoy
(0) = (e, — wy) 3 + (7.290)
A graph of 4(v) is shown for c, — vg > 0 in Fig. 7.28. A similar graph of the
pseudopotential (vu) for (¢, — vo) < 0 is shown in Fig. 7.29. Only the second
form is suitable for our purposes. This is because we desire a localized wave form,
u(x’ + ee) = 0. This will occur in Fig. 7.29 when the pseudoparticle leaves
v = Owhen the pseudotime x’ > —0°, falling once through the well to reach vxNonlinear lon-Acoustic Waves 175
(0)
Fig. 7.28 Sketch of pseudopotential when c, > vg
atx’ = 0, and taking an infinite amount of pseudotime x’ to fall back through the
well to reach v = 0 as the pseudotime x’ — +20, We thus obtain the shape shown
in Fig, 7.30. The pseudopotential in Fig. 7.28 would not allow Wx’ + +) — 0,
Let us now solve (7.287) exactly, with c, — vy < 0 or vy > ¢,, We all know
how to solve force equations of the form (7.287). Multiply (7.287) by v’ and
integrate, to obtain,
Stes
FOF = Qo — 6) Fo - (7.291)
where we have chosen the constant of integration to be zero because we want
v' = 0 when v = 0 (Fig. 7.30). Then
oye (2)" [ee -a - 4y" (7.292)
or
= (2) "ae (7.293)
[= a
Seep eaee es
Each side of (7.293) can be integrated. The left side is of the form
dv dv
7={—S— =; —2— 7.294)
Se = pv J w/I — Bo cae
where B = 1/[3(v9 — ¢,)]. Letu = VT — Bu, then v = (1 — w*)/B and du =
(-B dv/2)/\/1 — Bu. We find
d(o)
Fig, 7.29 Sketch of pseudopotential when c, < vy176 Fluid Equations
vO)
Bax
x
Fig. 7.30 Sketch of soliton solution to the Korteweg-deVries equation.
1 )= In (} = “) (7.295)
T+u t+
Damage geey = (2)"r
(42> ) » (4 +4) = =) x (7.296)
from (7.293). With y = [(v, — ¢,)/a]”, and exponentiating both sides, we get
i 7 7 =e (7.297)
Then | — u = (1 + w)e?”, implying that
le
u= ora (7.298)
and v = (1 — w)/Bis
_ 1 fa tery —U - ey
Bal | en (7.299)
or
eal 4 eee aaa
US GREE aay = | OD (7.300)
which is
ara
v = 3(v — ¢,) sech? (2) x | (7.301)
In fact, this solution has only been derived for x’ < 0 since we chose the v’ > 0
branch in (7.292); nevertheless, it would be easy to obtain the part of (7.301) for
x’ > 0 by choosing the v’ < 0 branch in (7.292); therefore, (7.301) applies to all
x’ and is the soliton solution. Note that the larger amplitude solitons are mote
sharply peaked, having a smaller scale length. This behavior is in accordance with
our picture of nonlinearity » a,v which balances dispersion 4,3 (Fig. 7.31). Back
in the lab frame, where x = x’ + vot, this solution is
= )e - wo | (7.302)
v(x,0) = 3(v9 — ¢,) sech? [(2a*Nonlinear Langmlur Waves 77
O(n’)
Large and narrow
Small and wide
e
Fig. 7.31 Soliton solutions exhibit the balance between dispersion and nonlinearity,
EXERCISE Write out sech? (x’) in terms of exponentials, and show that it
reproduces the solition behavior shown in Fig. 7.31.
7.16 NONLINEAR LANGMUIR WAVES—
ZAKHAROV EQUATIONS
In the previous section, the addition of one nonlinear term to the equation for
ion-acoustic waves led to a nonlinear wave equation with soliton solutions. In this
section, the addition of a different nonlinear term, representing the ponderomotive
force, to the equation for ion-acoustic waves leads to a set of coupled nonlinear
wave equations that describe the nonlinear interaction between high frequency
Langmuir waves and low frequency ion-acoustic waves.
Consider a collection of linear Langmuir waves in one spatial dimension whose
electric field (the subscript A stands for high frequency) can be written
Ena) = > E(x) exp(-ia,t) + c-. (7.303)
The amplitude £(x,1) contains the Ak7AZw, frequency dependence of the Lang-
muir waves. Since K?A2 << 1 for Langmuir waves, the function £(x,1) varies
slowly in time compared to the rapidly varying exp (—w,t). Thus, in the ponder-
omotive force equation (2.76) the constant field Ey can be replaced by the slowly
varying amplitude £(x,1) so that the low frequency ponderomotive force acting on
electrons is
a
wo eer (7,304)
4m.u,
where the plasma frequency «, appears in the denominator because all compo-
nents of the Langmuir wave field E, have frequency near the plasma frequency.
Our goal is to rederive the ion-acoustic wave equation including the pondero-
motive force (7.304). This couples the low frequency ion-acoustic waves to the
high frequency Langmuir waves, If we then rederive the high frequency Langmuir-
wave equation including the change in the background density due to the presence
of ion-acoustic waves, we will have two coupled nonlinear equations in the two
unknowns representing Langmuir-wave electric field and ion-acoustic wave densi-
ty perturbation.178 Fiuld Equations
The derivation of these equations makes explicit use of the fact that the problem
has two time scales. Thus, we shall often encounter equations of the general form
a(t) + b(t) exp (—iw,t) = e(t) + d(1) exp (—ia,1) (7.305)
where a, b, c, and d vary slowly compared to the time scale w,"', that is,
1 da 1
az “2 a < 0 everywhere, (7.339) can be integrated to yield
Join (L=*) = ge
2 in (+ ++) O22 (7.340)
Solving for u and converting back to f, one finds
f= (20)!? sech (0'72) (7.341)
so that the total field, as given by (7.333), is
E(z,r) = (20)'* exp (iMr) sech (02) (7.342)
which can be called a Langmuir soliton.
EXERCISE Sketch the solution (7.342) and show that it is indeed a localized
“bump.” Sketch the density perturbation (7.331).
A more general class of solitons exists [23], moving at any speed the absolute value
of which is less than the sound speed.
In the next section, we turn our attention to another important subject that can
be studied within the context of the Zakharov equations: parametric instabilities.
The study of solitons and parametric instabilities is one of the most active areas of
research in plasma physics [24-26].
7.17 PARAMETRIC INSTABILITIES
Consider a plasma that contains a single plane wave of finite amplitude. Within
the fluid theory, the system of plasma plus wave can be thought of as a time-
dependent equilibrium state. We can then ask the question: Is such an equilibrium
stable or unstable? This is the same question we asked about time-independent
equilibria in Chapter 6 on Vlasov theory and in Section 7.13 on the two-strcam
instability. The answer to the question often indicates instability, and such insta-
bilities are called parametric instabilities, the “parameter being the amplitude of
the single wave.182 Fluid Equations
One can look for such instabilities with any of the waves studied in this book.
For example, we shall use Langmuir waves, the stability of which can be studied
within the context of the Zakharov equations of the previous section. It turns out
that the most general instability in this case involves the single finite-amplitude
Langmuir wave, two other Langmuir waves, and one low frequency wave. The
stability analysis proceeds by assuming that the amplitudes of the two other
Langmuir waves and the low frequency wave are infinitesimal. We choose
E(z,t) = Ey exp (—iwpt + ikoz) + E, exp [—i(a) + w)r
+ i(ky +'k)z] + E. exp [iw ~ w)r + (ky — k)z) (7.343)
and
n = fiexp(—iwr + ikz) + complex conjugate (7.344)
where fi, E_, and E, are all much smajler than Ey. The equilibrium solution
E(z,t) = Eo exp (—fwot + ikoz), n(z,7) = 0, is chosen to satisfy the Zakharov
equations with Ey real.
EXERCISE Show that this solution implies wy = k,?.
Inserting the forms (7.343) and (7.344) into the first Zakharov equation (7.329),
and keeping only those terms with spatial dependence ~ exp(i{ky + k)z), we find
{wy + WE, — (ko + ky Ey = AEo (7.345)
Likewise, the terms with spatial dependence ~ expli(ky — k)z] yield
(w) — w*)E. — (ky — KE. = fi*Ey (7.346)
Solving (7.345) and (7.346) for E, and £_, inserting these into the second Zakharov
equation (7.330), keeping only terms with spatial variation ~ exp (ikz), and elim-
inating fi from each term yield the dispersion relation
OR = REP (s — met =e =) sr sek) (7.347)
There are several types of solutions. With ka > 0, we first look for an instability
with k < 0. If |w| is small, the second denominator on the right is larger than the
first, so we ignore the second. This is equivalent to ignoring the term E_ in the
electric field (7.343), so this instability involves only Ey, E., and fi and is thus
known as a three-wave interaction. The dispersion relation is now
(a? — kK) (@ — 2kky — Ke) ~ REZ = (7.348)
Looking for a solution with w = k + éwhere |8| << |k], we write (w? — k’) =
(w + kw — k) = (2k + 8)8 ~ 26; Eq. (7.348) then yields
2
8 + Ok — 2kky — kK) — bi =0 (7.349)
At the particular negative wave number that satisfies k — 2kk, — k? = 0 or
k= — 2ky + I, this is183
Parametric Instabili
é6= (7.350)
which indicates instability since k < 0. If ky >> 1, this becomes
5 = ikp!?Ey : (7.351)
EXERCISE Show that in physical units denoted by a tilde, ky >> 1 means
Ryd. >> (m,/m)'?.
EXERCISE What does |6| << |k| mean in physical units?
Since the electric field £, has a wave number ky + k = ky — 2ky + 1 = — ky +
1, £, has a negative wave number and travels in the opposite direction to Ep. It is
known as a backscatter instability, and is one example of a parametric decay
instability. The physical growth rate 4 is
% _ [mn ;
on te) (Kodo)?
Wan,Fy™ (7.352)
where physical quantities are denoted by a tilde.
EXERCISE Demonstrate (7.352) from (7.351).
The dispersion relation (7.347) also yields an instability that involves all of the
terms and thus is known as a four-wave interaction. The simplest case is when
wo = ko = 0; that is, the physical field represented by By is oscillating exactly at
the plasma frequency w, and has zero wave number (a so-called dipole field).
Looking for a purely growing instability w = iy, we see that the dispersion
relation (7.347) becomes
OP + RG? + KY) — 2k EB? = 0 (7.353)
the solution of which is
- + (+ kt) + + [2 — KAP + BEET? (7.354)
2
¥
With both k << Land £, << 1 the k* term within the bracket can be discarded,
and the square root can be expanded to yield
y = K(QE? — k?)"? (7.355)
which is the growth rate of the four-wave interaction known as the oscillating
two-stream instability [27]. The growth rate versus wave number is sketched in Fig.
7,32. The maximum growth rate -y = E,? occurs atk = +E.
EXERCISE Show that in physical units these are
E,
Saat (7.356)
V/a, =
and
w= (afi)
ane (7.357)184
~2 5,
Fiuld Equations
7
Ey E, 2B,
Fig. 7.32 Growth rate versus wave number for the oscillating two-stream instability.
The study of parametric instabilities is very important for such fields as lager
fusion, particle beam fusion, radio-frequency heating of the ionosphere and of
magnetic confinement devices, and solar radio physics.
This brings us to the end of our study of the fluid equations of plasma physics.
In the next chapter, the fluid equations for each species are combined to yield the
equations of magnetohydrodynamics.
REFERENCES
mH
(2)
3]
{4]
[5]
[6]
7
[8]
9]
[10}
a
12
113}
14}
us)
K. R. Symon, Mechanics, Addison-Wesley, Reading, Mass., 1960.
A. N, Kaufman, in Plasma Physics in Theory and Application, edited by
W. B. Kunkel, McGraw-Hill, New York, 1966, p. 91.
N. A. Krall and A. W. Trivelpiece, Principles of Plasma Physics, McGraw-
Hill, New York, 1973.
P. C. Clemmow and R. F. Mullaly, in Physics of the Ionosphere: Report of
Phys. Soc. Conf. Cavendish Lab., Physical Society, London, 1955, p. 340.
W. P. Allis, in Sherwood Conf. Contr. Fusion, Gatlinburg, Tennessce, April
27-28, 1959, TID-7582, p. 32.
F. F. Chen, Introduction to Plasma Physics, Plenum, New York, 1974.
T. H. Stix, The Theory of Plasma Waves, McGraw-Hill, New York, 1962.
W. M. Manheimer, da Introduction to Trapped-Particle Instability in Toka-
maks, TID-27157, National Technical Information Service, Springfield, Vir-
ginia, 1977.
N. A. Krall, in Advances in Plasma Physics, Vol. 1, edited by A. Simon and
W. B. Thompson, Wiley-Interscience, New York, 1968, p. 153.
T.K. Chu, B. Coppi, H. W. Hendel, and F. W. Perkins, Phys. Fluids, 12, 203
(1969).
B. Coppi, Revista de? Nuovo Cimento, 1, 357 (1969).
P. Rutherford and E. Frieman, Phys. Fluids, 11, 369 (1965).
A. A. Rukhadze and V. P. Silin, Sov. Phys, Uspekhi, 11, 659 (1969).
J. D. Jukes, Rep. Prog. Phys., 31, 305 (1968)
P. H. Rebut, Plasma Phys., 9, 671 (1967).Problems 185
[16] A. B. Mikhailovskii, Rev. Plasma Phys., 3, 159 (1967).
[17] B. B. Kadomtsev and O. P. Pogutse, Rev. Plasma Phys., 5, 249 (1970).
[18] A. B. Mikhailovskii, Theory of Plasma Instabilities, Vol. 2: Instabilities of an
Inhomogeneous Plasma, Consultant Bureau, New York, 1974,
[19] F. F. Chen, Sei. Am., 217, 76 (1967).
[20] D. J. Korteweg and G. deVries, Phil. Mag., 39, 422 (1895). ;
[21) V. E. Zakharov, Zh. Eksp. Teor. Fiz., 62, 1745 (1972) [Sov. Phys.-JETP, 35,
908 (1972)].
[22] P. J. Hansen and D. R. Nicholson, Am. J. Phys., 47, 769 (1979).
[23] G. Schmidt, Phys. Rey. Lett., 34, 724 (1975).
[24] A.C. Scott, F. Y. F. Chu, and D. W. McLaughlin, Proc. IEEE, 61, 1443
(1973).
[25] G. B. Whitham, Linear and Nonlinear Waves, Wiley, New York, 1974.
[26] K. E. Lonngren and A. C. Scott, eds., Solitons in Action, Academic, New
York, 1978.
[27] _K. Nishikawa, J. Phys. Soc. Jpn., 24, 916, 1152 (1968).
PROBLEMS
7.1 Energy Transport Equation
Obtain an equation for the fluid transport of particle kinetic energy by multiplying
the Vlasov equation by ¥sm,v? and integrating over all velocity space. Simplify
your result in any convenient fashion.
7.2 Fluid Conservation Properties
Suppose we have an electron-proton plasma that is finite in extent in all three
dimensions. Suppose there is no magnetic field. Using the fluid equations, prove
that
(a) For each species, total particles are conserved.
(b) Momentum is not necessarily conserved for each species.
(c) Total momentum, summed over species, is conserved.
7.3. Langmuir Waves
One does not need to look for individual sinusoidal wave solutions to solve wave
equations. Consider the electron fluid equations in the differential form (7.38) to
(7.40).
(a) Combine these equations, and linearize, to obtain the linear partial differ-
ential wave equation
7? ~ w? — 3u? de)E(x.) = 0186 Fuld Equations
(b) Suppose the initial conditions for E(x,1) are
E(xt = 0) = f(x),
EG, = 0) = 0
where an overdot indicates a time derivative. Using Fourier and Laplace
transform techniques, find an exact explicit solution for the time evolu-
tion of E(x,1).
(c) Suppose f(x) represents a standard wave packet, a sinusoidal variation
with space accompanied by a Gaussian envelope
E(x, = 0) = f(x) = Eye sin kegx
where we assume ky >> L”!. By using an appropriate approximation, if
necessary, in the exact solution from (b), show that the wave packet travels
with the group speed
eo ee
Vel = | Gk [ant = (Koad
Show also that the packet spreads as it propagates, and the rate of spread-
ing (dispersion) is proportional to |dV,. SOK ak, Does the packet, move to
the right, to the left, or does it split into right- and left-going pieces?
7.4 Negative Energy Waves
Suppose a plasma has cold electrons drifting with velocity vp with respect to cold
ions. Derive the wave dispersion relation corresponding to high frequency elec-
tron plasma waves. Show that in the frame moving with the electrons, these are
just our old cold plasma waves. Plot the two branches of w(k) vs. k. Use the wave
energy formula (6.72) to evaluate the wave energy. Indicate the regions of your
dispersion diagram where the energy is negative.
7.5 Upper Hybrid vs. Right Cutoff Frequency
Prove that the right cutoff frequency
2
op = a + [er + (a24)]
is always greater than or equal to the upper hybiid frequency
a = (w2 + 02)8
7.6 Upper Hybrid Wave
Compare the derivations of the upper hybrid wave (a perpendicular, electrostatic
wave) and the extraordinary wave (a perpendicular, partially electromagnetic and
partially electrostatic wave). Does the assumed form of the upper hybrid wave
satisfy Maxwell’s equations? Why not? In which parameter regime does it approx-
imately satisfy Maxwell’s equations? In this parameter regime, is there any differ-
ence between the upper hybrid wave and the extraordinary wave? In the X-mode
derivation, show that Faraday’s law contains the same information as would be
contained by Poisson's equation plus the electron continuity equation. ReproduceProbiems 187
the graph of x = ck/w for the extraordinary wave and draw the dispersion dia-
gram for the upper hybrid wave. For which portion of each diagram do we trust
the derivation? Which portion of the X-mode diagram corresponds to the upper
hybrid wave?
7.7 Model of Collisions
We wish to derive a simple model of collisional effects in a plasma, Suppose that a
typical particle suffers collisions at a rate vy. Then the particle when oscillating in
the electric field of a wave will occasionally suffer a collision (assuming »y << w)
and to the first approximation can be thought to lose all of its directed energy. An
electron fluid element with velocity v will thus lose momentum at a rate —vmavm,,
where we have assumed that each colliding electron loses momentum ~m,v. Thus,
we can add a term in the electron force equation to represent collisions,
nm, OV, + nmV,+ UV, = — UP, — en,E — vnm,V,
With this extra term, rederive the Langmuir wave dispersion relation, assuming
v << w. At what rate does the electric field damp away? At what rate does the
wave energy damp away?
7.8 Low Frequency Dielectric Constant
Recall that in the theory of dielectrics, one likes to include the currents in
_ 4a a
VX B= J+ 4k
in the dielectric function ¢; thus
el
vx B=~ aD
L—_——_—_——_—>
Fig. 7.33 Configuration for Problem 7.9.
where188 Fluld Equations
D=c«
Suppose that a slowly varying sinusoidal electric field is applied across a magnetic
field. Derive an expression for ¢ by considering the polarization current produced
by the electric field. What do you suppose “‘slowly varying" means?
7.9 Kunkel's Problem
A plasma of mass density p = 1o(m, + m,) is bounded by two parallel conduct-
ing plates separated by a distance L. A gravitational acceleration g is applied at
right angles to a uniform magnetic field B, and both of these are parallel to the
plates as shown in Fig. 7.33. Show by means of a careful particle drift analysis that
the plasma can accelerate freely downward only if switch S is open, and if the low
frequency dielectric function from the previous problem is € >> 1, What is the
voltage between the two plates in that case? If 5 is closed, what is the current
density between the two plates?
7.40 Laser Fusion
In order to obtain controlled thermonuclear fusion using deuterium and tritium,
one needs to satisfy the Lawson criterion nr > 10'* (c.g.s.) at a temperature
T ~ 10 keV, where n is the number of particles per cm’, and 7 is the confinement
time in sec. Use the Lawson criterion to derive the corresponding requirement for
laser pellet fusion, pr > 1 (c.g.s.) where p is the density of the compressed pellet
in g/cm? and r is the radius of the compressed pellet in cm. (Hint: How does one
define “confinement time” for inertial ‘confinement’? How can one estimate this
physically?)CHAPTER
Magnetohydrodynamics
8.1 INTRODUCTION
Chapter 7 is concerned with a set of equations, the fluid equations, which were
derived from the Vlasov equation, which in turn was derived from the Klimonto-
vich equation by neglecting all collision effects. Thus, all of the phenomena dis-
cussed in Chapter 7 will occur only when collisions are not important, A rough
criterion for the importance of collisions is obtained by comparing the collision
frequency v,; to the frequency w of the phenomenon under consideration; the fluid
treatment is valid when »,, << w. Since we have seen (Section 1.6) that the
collision frequency vq = w,/A, there is a huge range of frequencies where the
fluid treatment applies. However, there is a significant range of frequencies, 0 S
@ < v,, where the fluid treatment does not apply. In particular, one often wants
to find an equilibrium plasma configuration; for example, in tokamaks, mirror
machines, planetary magnetospheres, pulsar magnetospheres, and stellar winds.
An equilibrium is equivalent to w = 0(d, = 0), and collisions must be included
in such considerations.
Let us then develop a set of equations that are valid for low frequencies. [Stu-
dents who have not studied Chapters 3 and 6 can skip directly to (8.3) and (8.7)
witb the understanding that the extra terms represent the physics of collisions.)
Recall the plasma kinetic equation (3.26),
as eae .
af tv VS + (@ He vxB) Wh
eae ae 7
ac ((oe + <% we) v.9N,} (8.1)190 Magnetohydrodynamics
Here, f,(x,v,/) is a smooth function obtained by averaging the Klimontovich densi-
ty N,(x,v.t) over an appropriate volume, while 6N, is the difference between the
smooth function f, and N,, which is the sum of delta functions. In Chapter 3 we
argued that the right side represents discrete particle effects, including collisions.
We can therefore invent a symbol for the right side of (8.1), (df/d1)., and write
(8.1) as
antevas+& (e+txs)-we= (4) ea
Thus, we are thinking of f, as the number of particles in a small volume of
six-dimensional phase space (x,¥), divided by that volume. (See the discussion in
Section 6.1.) Then (@ f,/0!), is the rate that particles are gained or lost by that
small volume because of collisions. Recall that E and B in (8.2) are also averaged
quantities, so that they do not include the fields due to individual particles. This
identification of (0 f,/d1), is admittedly crude, but we shall not attempt to do
better here. There does exist a large body of more exact literature that starts from
the formally exact expression (8.1). Here, we use (8.2) to try to obtain the most
significant effects of collisions
Having identified (0 f,/d1), as the change in f,(x,v,t) due to collisions, we expect
it to have a much stronger influence on the velocity dependence of f, than on the
spatial dependence of f,. This is because a collision can cause a huge change in a
particle’s velocity, but does not cause much change at all in a particle’s position.
In Section 7.2 we obtained the fluid equations by integrating the Vlasov equa-
tion (6.5) over velocity space after multiplying by an appropriate power of veloci-
ty. Let us repeat that procedure with (8.2). Multiplying by unity and integrating
over alt velocity space, we obtain
Aung (ut) + V-(Van,) = fede fy (8.3)
c
where the left side is as in (7,15). The right side represents the change in the
number of particles in a smal} volume of real space due to collisions; we have just
argued that this is very small since collisions do not cause large changes in particle
positions; therefore we set this to zero and obtain
Gn, (Kt) + Ve (n,V,) = 0 (8.4)
which is just the continuity equation (7.15).
Next, multiply (8.2) by v and integrate over all velocity space; we obtain the
force equation (7.28) with the addition of one term. This is
m,n,a,V, + mndV,+ VV, = — VP, + q.ndE + + Vv, x B)
a
+ m.faev (4). (8.5)
The term
(8.6)Introduction 191
represents the change in the momentum of species s at position x due to collisions.
A species cannot change its own momentum by colliding with itself; the center of
mass of two electrons, for example, is not accelerated during a collision of the two
electrons (see Section 2.9), However, the momentum of electrons in a certain
volume of space can certainly be changed by collisions with the ions. For example,
a beam of electrons incident on a plasma will slow down due to collisions with the
ions; the ions begin to move in the initial direction of the electron beam because
they have taken up the electron momentum. Thus, we expect K,(x) = — K,(x). It
would be possible to develop simple but crude models for K(x), but here we shall
teave K, in general form. Our force equation then is written as
mznz3,V, + mn(V."V)V, = — VP. + gn(E + a Vv, x B) + Kx) (8.7)
The fluid equations (8.4) and (8.7) are the same as the fluid equations in Chapter
7, without the term K,(x) in (8.7). When written for both electrons and for ions,
this set is called the two-fluid model. We now wish to combine the electron equa-
tions with the ion equations to obtain a one-fluid model, also known as the equa-
tions of magnetohydrodynamics (MHD). We thus wish to think of a single fluid
characterized by a mass density
Prdx) = mn(x) + mnlx) = mndx (8.8)
a charge density
PAX) = gndx) + gankx) = e(n; — 2) (8.9)
a center of mass fluid flow velocity
v= (may, + man¥.) (8.10)
Pu
a current density
JS gay, t+ anv, (8.11)
and a total pressure
P=P,+P, (8.12)
We wish to derive four equations relating these quantities: a mass conservation
equation, a charge conservation equation, a momentum equation, and a generalized
Ohm's law.
First we derive a mass conservation law. Multiply the ion continuity equation
(8.4) by m,, the electron continuity equation (8.4) by m,, and add to obtain
Y= (enV) (8.13)
which is the mass conservation law.
Next, multiply the ion continuity equation (8.4) by g;, the electron continuity
equation (8.4) by g,, and add to obtain192 Magnetohydrodynamics
8p. j=
et V=0 (8.14)
which is the charge continuity equation or charge conservation law.
Consider next the force equation (8.7). Regarding V, and d,n, as small quanti-
ties, neglecting the products of small quantities, and recalling K, = — K,, we add
(8.7) for electrons and ions to obtain
av
Pw ‘at
=~ w+ pkttixe (8.15)
which is the one-fluid force equation, or momentum equation.
Finally, we desire an equation for the time derivative of the current, called a
generalized Ohm's law. Multiplying the force equation (8.7) by 9,/m,, adding the
ion version to the electron version, neglecting quadratic terms in the small quanti-
ties dn, and V,, and using g, = — 9, = e, we find
‘
ve + 5 UP, + (Se + fn) p+ Sty, xB
2,
+fuvxp+(£+)x, (8.16)
me my m,
We notice that
ne e
me Meee ace ee ae eV me (mnN))
= I+ ee (mn, + mn.)
arr) vt meme (ou) (8.17)
where the first line merely adds and subtracts the same quantity, and the second
line adds the tiny quantity (m,/m,)V,, which is negligible compared to V, as
already incorporated into J.
Using (8.17) in (8.16), neglecting m;! << _m,”’ wherever possible, and assuming
that P, ~ P; ~ 4P and n, ~ n,, we find
Lai eeu i ee oe
ie? e+ SE +t Vv x B) FIX B+ TK B18)
Recall that K, represents the change in ion momentum due to collisions with
electrons. It is reasonable then to assume that K, is a function of the relative
velocity V, — V, between the two species; keeping only the first term in a Taylor
expansion of K,, we find
K; = CV, — V.)
= C3Introduction 193
=- fey (8.19)
mo
where the constant C, bas been put in a form such that o can be identified as a
conductivity, as we shall see. The minus sign has been chosen because we expect
collisions to decrease the current caused by relative species velocity. When we
multiply by m,,/pye?, (8.18) becomes
mem, 93 _ opi ppt yxp—-—
J
= xB
Pue? at 2pme ¢ Pye ae o
(8.20)
which is the generalized Ohm's law. The name comes from the fact that if the only
important terms are the second and the fifth terms on the right side, we have
J=oE (8.21)
which is Ohm’s‘law and in which o is clearly the conductivity.
This completes our derivation of the MHD equations, Collecting these equa-
tions, we have
a
Ht + Ve(euY) = 0 (8.13)
8p. =
a +V°S=0 (8.14)
ov a
pu Gp =~ VP +E + TIXB (8.15)
mem, dS _ im;
preuien Ge eueece
+tvxp-—jxp-i (8.20)
c Puree o
When coupled to Maxwell’s equations
1 aB
xpa- 1 2B
VXE ear (8.22)
and
daa elk oe
VXB= I+ a (8.23)
we have 14 equations in the 14 unknowns py, p., V, J, E, and B; this assumes that
the pressure P can be expressed in terms of the mass density py.
For very low frequencies, one can ignore the 4,J term in the generalized Ohm's
law, whereas for low temperatures the VP term can be ignored. In addition, when
the current is small we can neglect the J X B term (known as the Hall term)
compared to the V X B term; under al! these assumptions, Ohm’s law (8.20)
becomes
oe 1 He Jt
OS E+ TVXB- > (8.24)194 Magnetohydrodynamics
or
s=o(e+tvxs) (8.25)
When collisions vanish, the conductivity becomes infinite and, in order to have
only finite currents, we must have
E+ivxp=o0 (8.26)
or
E -ivxe (8.27)
Under this infinite conductivity, low frequency condition, no charge imbalances
are allowed and we have p, = 6. Under these ideal MHD conditions, our basic
equations (8.13), (8.14), (8.15), (8.20), (8.22), and (8.23) become
4.py + V+ (pyV) = 0 (8.28)
pyOV =~ VP + 4 JxXB (8.29)
Vx (Vx B) = aB (8.30)
vxp=“Zy (8.31)
where the low frequency assumption is used to ignore (1/¢)(E/dr) on the right of
(8.31).
We shall not attempt to further justify (8.28) to (8.31); instead, we shall take for
granted that these equations can be justified in useful physical situations. In the
next section, we shall use these equations to consider the equilibrium and stability
of various plasma configurations.
8.2. MHD EQUILIBRIUM
In many cases one is interested in the equilibrium configurations of a plasma. Once
the equitibrium is found, one then asks whether or not the equilibrium is stable.
For example, the problem of the earth’s magnetosphere and its interaction with
the solar wind can be approached by first tooking for MHD equilibria. (Since
cauitibrium implies that all zeroth order quantities have no time derivatives, we
can feel some confidence in using the ideal MHD equations to look for equilibria.)
Once we find the zeroth order quantities B°, V°, J°, and p,,° (and thus P*) satisfying
the ideal MHD equations (8.28)-(8.31) with 8, — 0, we can then linearize about
the equilibrium to determine whether it is stable. In other words, we let py, =
Pa? + py, etc. and 4, — — iw, and we solve for all possible values of w. If one
of these values of w has Im(w) > 0, we have instability and any tiny pérturbation
of the equilibrium will grow as exp [Im(w)¢] until the equilibrium is destroyed. If
no unstable values of w are found, then we have MHD stability, but this does not
imply overall stability. Much of the high frequency physics has been lost in firstMHD Equilibrium 195
going from the Vlasov equation to the two-fluid equations, and in next adding the
two-fluid equations to obtain the MHD equations. A system that is MHD stable
may welt be unstable when two-fluid effects or Vlasov effects are considered. Thus,
MHD stability is a necessary but not a sufficient condition for overall stability.
Before proceeding to questions of MHD equilibrium and stability, let us try to
gain a measure of intuition regarding our MHD equations. From Ohm's law
(8.25) and Maxwell’s equations (neglecting the displacement current)
vxp=4ty (8.32)
and
vxe=—tap (8.33)
we obtain
vx qv xB) = 22 [ + +lvxwxB (8.34)
or
aB= 0 x (Vx B)+ & vB (8.35)
Thus, the magnetic field at a point in a plasma can be changed by the fluid
convection, the first term on the right side of (8.35), or by diffusion due to the
second term on the right side of (8.35). When V = 0, (8.35) is
=f
0B = 7 VB (8.36)
which is the standard form of a diffusion equation; the constant c?/da is known
as the magnetic diffusivity, Dimensional analysis of (8.36) shows that the field can
diffuse away with a diffusion time 7p given by
2
ty ~ “ESE (8.37)
¢
where L is the scale length. Since the conductivity o is inversely proportional to the
collision rate, the physics of the diffusion must involve the disruption of particle
orbits due to collisions, which in turn disrupts thé’current that gives rise to the
magnetic fields; the disruption of the current allows the magnetic field to diffuse.
‘We consider next a very important concept, that of frozen-in field lines. Consid-
er a surface AS drawn perpendicular to the field lines, and suppose that the
boundary of this surface is moving with some specified velocity field (Fig. 8.1)
What is the time rate of change of the total magnetic flux
® = fia “B (8.38)
through the surface AS? It is
b=f aa-b+f di-s (8.39)
a a196 Magnetohydrodynamics
Fig. 8.1 Surface AS perpendicular to the magnetic field lines.
where the first term gives the contribution from the time rate of change of B, and
the second term gives the contribution from the movement of the boundary of the
surface AS, At every point on the boundary of AS, the rate of change of the area
enclosed by the boundary is proportional to the perpendicular component of V at
that point, and therefore it is given by V X dl. Thus, (8.39) is
b = . B :
é =J.aa a +B x ap (8.40)
In ideal MHD (a ~ 0) we define the surface AS to be attached to the fluid, and
we evaluate (8.40) for &. We first integrate (8.30) over the surface AS, obtaining
f dA-aB f dA+[V x (VXB)] (8.41)
as as
If we recall Stoke’s theorem,
fx Orda =pera (8.42)
as
(8.41) becomes
f dA- 0B =fa -(V x B) (8.43)
as
Next recalling the vector identity
A+(B XC) =(A XB)-C (8.44)
we find
f dA: 0B +f$B-wW x ay =o (8.45)
as
But by (8.40), the left side of (8.45) is just the total time rate of change of ®
through the surface AS; therefore
é=0 (8.46)
as long as the surface AS moves with the fluid.
Suppose that at ¢ = 0 we draw two surfaces AS, and 4S, and we then sweep
the surfaces along the instantaneous field lines to form two flux tubes, as shown in
Fig. 8.2. If the two surfaces AS, and AS, intersect at one point, then there is oneMHD Equilibrium 197
AS,
Special
field
line
AS,
Fig. 8.2 Two flux tubes.
special field line that is the line at which the two flux tubes touch. By coloring the
fluid in one flux tube red and the fluid in the other tube blue, we can follow the
two flux tubes forever. For any reasonable flow, the tubes will always touch, and
the line of touching identifies that particular magnetic field line forever. Thus, in
ideal MHD, we can similarly label each and every field line, and we can say that
the plasma is frozen 10 the field lines.
In nonideal MHD (0 # &), it becomes much more difficult to label field lines.
This is true partially because lines can disappear because of resistivity.
Let us now return to consider the requirements for an MHD equilibrium.
Looking for equilibrium solutions to (8.28)-(8.31), with no fluid flow, we require
that
o=-vwPe+t sxe (8.47)
and ‘
an
vxB=y (8.48)
which yield
1 a
VP - (Vv xB) x B=0 (8.49)
Recalling
(Vx B) x B= (B-WyB- 4 vB (8.50)
we have
1 = g.
vp + ve = weve (6.51)
When (B + V)B = 0, this is
(8.52)
which leads us to define the magnetic pressure B’/87, and to state that in equilib-
rium, magnetic pressure must balance plasma pressure (Fig. 8.3),
B
ee ‘eae constant (8.53)
Returning to (8.47), we see that in equilibrium,198 Magnetohydrodynamics
Fig. 8.3 Magnetic field pressure balances plasma pressure in ideal magnetohydrodynamic
equilibrium.
w=tixe (8.54)
so that B: VP = J+ VP = 0; in other words, B and J lie along surfaces of
constant pressure.
Let us consider two simple cases of plasma equilibria, the theta pinch and the
z-pinch, In the theta pinch, capacitor plates are discharged about a cylindrical
conductor, as shown in Fig. 8.4. The azimuthal J makes a B into the paper; the B
into the paper induces an azimuthal E in the direction opposite to J, which in turn
produces an internal current in the plasma in the direction opposite to J. A
hypothetical final state could be as shown in Fig. 8.5 where the central plasma
region has B = 0. The pressures are’ then as shown in Fig. 8.6 so that P +
B’/8 = constant, and there is an azimuthal current sheet at rp such that
ve=tixs (8.55)
is satisfied (Fig. 8.7).
A second possible equilibria is the z-pinch, where a current flows along a plas-
ma, and the plasma is confined by its own magnetic field and its own current
Conductor.
Fig. 8.4 Theta pinchMHD Equilibrium 199
Fig. 8.5 Hypothetical final state of a theta pinch.
through the J XB force, as shown in Fig. 8.8. Then in cyclindrical coordinates we
have
vPp=1yxp= 1 (vxp)xB (8.56)
ie an
or
oP B, a
Peet tH eee
ar tar ar ("Be) (8.57)
The solution of (8.57) will be left for one of the problems; here we merely note that
(8.57) does have well-balanced solutions for B,(r) and P(r).
Before closing this discussion of equilibrium, we note two related points. First,
because
P + = = const; 8.58
3 instant (8.58)
the magnetic field is Smarr inside 2 plasma (P > 0) than outside a plasma
(P = 0). This is another illustfation of the fact that a plasma is diamagnetic; we
have seen this before i considering singte particle motion.
Second, we notice from (8.48) and (8.49) that when J is parallel to B, we have
vP=0 (8.59)
and
(Vv x B) XB=0 (8.60)
Pressures
Magnetic pressure
Radius
Fig. 8.6 Pressure balance for the final state of a theta pinch shown in Fig. 8.5.200 Magnetohydrodynamics
Current
density
Radius
%
Fig. 8.7 Current profile that maintains the pressure profile in Fig. 8.6.
which is known as the force free situation. Any magnetic field for which
VX Bix) = f(x) B(x) (8.61)
satisfies (8.60).
In the next section we proceed to discuss the stability of our equilibria. It is the
unfortunate case that both the theta pinch and z-pinch are unstable, as well as the
simple mirror machine. We shall find that stable equilibria are possible, but only
when certain criteria are satisfied.
8.3 MHD STABILITY
In the last section we found various examples of MHD equilibria. We must now
ask whether those equilibria are stable. There are two ways of doing this. The first
is to linearize the equations of motion about the zero-order equilibrium, and solve
for the freaicaete neatly ee same way in Which we previously found linear
waves. If one of these frequencies has Im(w) >_0, then exp(—iwt) ~ exp{Im(w)#)
will grow with time, and the system is unstabie. The second is to consider the total
energy of a system, and (Sask whether har energy increases or decreases under a
perturbation. If the energy increases, the perturbation will not grow. If the
energy decreases, the perturbation can happen and have energy left over to go into
kinetic energy of expansion; this is the mark of instability.
The first method is to linearize the equations of motion. In ideal MHD, these
are (8.28) to (8.31). For example, consider a plasma held up against the force of
ssid
B
Fig. 8.8 Current and magnetic field configurations in a z-pinch.MHD Stability 201
Ci
y
© 0 0
BO a
Fig. 8.9 Plasma held up against the force of gravity by a current sheet.
gravity by a magnetic field, as in Fig. 8.9. For this example, because we generalize
the force equation (8.29) to include gravity, we have
pudV =~ VP + + IX Bt pve (8.62)
Since |B| has a discontinuity at z = 0, (8.31) predicts a sheet current at 2 = 0 in
the (~)f-direction. The equilibrium quantities are, generalizing (8.51) and (8.52),
a BY _
77 t+ en — puk (8.63)
We next consider a perturbed fluid velocity at the plasma-vacuum interface with
sinusoidal variation only along y, an undetermined variation in z, and no variation
in x,
V = v(z) exp (—iwt + iky) (8.64)
where v = (0,v,,u,). It can then be shown [1] that instability results, with
a? = ~ kg (8.65)
or
w = + i(kgy'? (8.66)
which implies instability. This is called the Kruskal-Schwarzchild instability, and is
the MHD analog of the fluid Rayleigh-Taylor instability.
It is interesting to consider the microscopic physics of this instability. Recall the
current in the ~f-direction. Microscopically, this current is due to the g X B drifts
of the particles on the plasma surface. Since this drift is proportional to mass
(why?), the ions are drifting much faster. Now consider the initial perturbation, as
shown in Fig. 8.10. Since the ions drift to the left, and the electrons drift to the
right, charges build up as shown. This creates an electric field as shown. The
plasma on the surface then performs anE X B drift, down in the left section and
® + © +f ®
Fig. 8.10 Microscopic picture of the Kruskal-Schwarzchild instability.202 Magnetohydrodynamics.
Flute
a :
i}
m=1 me3
Fig. 8.11 Azimuthal variations of displacement for m = 0, m = 1, and m = 3.
up in the right section, thus intensifying the initial perturbation and leading to
instability.
A similar analysis could be carried out for other equilibria, such as the theta
pinch and z-pinch equilibria considered in the previous section. Consider the
z-pinch equilibrium, and perturbations ~ exp (—iwf + im@ + ikz), where 6 ts
the azimuthal angle (Fig. 8.11). Form = 0, the instability is known as the sausage
instability (Fig. 8.12). For m = 1, the instability is known as the kink instability.
Higher values of m are known as flute instabilities, because their perturbations
resemble fluted Greek columns.
We come to the second method of treating the question of stability in MHD
systems, the energy principle. Consider a ball in a potential well, as shown in Fig.
8.13. In the unstable case, a small change in the particle’s position leads to a
decrease in the particle’s potential energy; the difference in energy is available for
kinetic energy and the implication is instability. On the other hand, the stable case
is characterized by a positive change in potential energy for a small perturbation;
thus, the perturbation is prohibited and the system is stable. It is interesting to
contemplate the modification of these ideas when nonlinear effects are included.
It turns out that plasma systems behave in the same way. The energy of a
plasma, which potentially could be turned into the kinetic energy of instability, is
the integral over its volume of B?/87, the magnetic energy, plus its internal kinetic
energy “AT per particle (J, = 7; = T). Thus, the plasma energy W is
ae Bs
w = fw a+ 4 favor (8.67)
or
w= fa (= +2 3 P) (8.68)
If a hypothetical perturbation causes a decrease in W, the system is unstable. (We
will not prove this here.) If W increases, the system is stable to that perturbation.
Serre ee rere eee
Te
Fig. 8.12 Spatial variation of displacement for an m = 0 instability.MHD Stabillty 203
Ye in
(a) 7) @ (a)
Fig. 8.13 Examples of stability: (a) unstable; (6) stable; (c) linearly stable, nonlinearly
unstable; (d) linearly unstable, nonlinearly stable
(Note Murphy’s eighth law: To prove instability, one needs to find only one
unstable perturbation. To prove stability, one needs to prove stability for each and
every possible perturbation.)
One could apply this technique to each of the equilibria considered previously;
each would yield a change 6W < 0 for the types of perturbations considered
before. Here we consider the more general problem of the hypothetical inter-
change of two neighboring tubes of magnetic flux. If this teads to 3W <0, we will
call it an interchange instability (Fig. 8.14). We consider separately the change of
magnetic energy W,,,
Rr
Ww, = fav & (8.69)
and the change of internal plasma energy W,,,
Ww, = 3 fave (8.70)
The interchange is accomplished by moving tube ® to where @ used to be. Then
% BP
We = %, faa, Fo 7
x | alias ee (8.71)
where A, is the cross-sectional area, and /; is the length, of the ith flux tube. But ina
flux tube the flux is constant; therefore y, B,A,, or
«df
Wem Bae Ls
The change 6W,, in W,, by moving @ to ® is [@ keeps its flux but finds a new f,
and A,]
(8.72)
®
©
a
Fig. 8.14 Two neighboring flux tubes.204 Magnetohydrodynamics
(Woe = 73)
while the change for @ — Ois
i dl;
(6W,,Jo-@ = Sad, Ay (8.74)
The total change, adding (8.73) to (8.74), is thea
aw, = (22 = set] 4 ine -f £4] (8.75)
0 A, id
or
_ _ ate’) We
aw, = ~ 424 5 f + | (8.76)
If we pick two flux tubes with equal flux, we have
aw, = 0 (877)
‘With two flux tubes of equal flux, we next calculate the change in internal energy
of the flux tubes, as the plasma expands or contracts to fit into a changing volume
(Fig. 8.15)
The change in internal energy as we move the plasma in V, to the volume V; is
{assuming |V, — Vi|/V, << 1)
ee ceed Pyy
dW, o.@ = HPV =F 8 i)
2 (PV), 6P-y
Serna — prover
+ (PV), — yh, — Vi) (8.78)
\
Fig, 8.18 Neighboring flux tubes with equal flux may have different areas and volumes.MHD Stability 205
where we have approximated V = V, ~ V; in the denominator only, and we
have used the fact that PVY = constant in an adiabatic compression; assuming a
three-dimensional compression we have y = 5/3.
Next, the change 6, @_@ is obtained from (8.78) by interchanging © and :
adding the two pieces yields
__3 eV,
ow, =- 0 - y [GO - SJ aw - (8.79)
Approximating V, = V; in the denominator only, we have
eH Se CD Hontty):
Wy = — > ay PY NEY (8.80)
where 6(PV) now means (PV); — (PV). With y = 5/3 this is
6W, = Vi~v8(PV)SV (8.81)
or
8W, = Vi-YSVSPV,Y + yP,V7-'3V) (8.82)
Now suppose we are in a low density part of the plasma, so that P(6V/V) << 6P.
(This is possible because 6V will be very small when magnetic fields are large;
recall that the flux of the two tubes is equal and we only need a small magnetic
field gradient VB?/81 ~ BVB to balance the particle pressure gradient VP). Then
OW, = 6V bP (8.83)
If the plasma density decreases from 1 to 2, 6P < 0, there will be instability
(8W, < 0) if 6V > 0 or
av = sf did = 08 oo (8.84)
so that the condition for instability is
D) (©
D) (CO
Fig. 8.16 Magnetic field lines and plasma density in a cusp configuration.
Plasma206 Magnetohydrodynamics
=o
pny mina
reer
So)
Fig. 8.17 Magnetic field curvature in present-day mirror machines.
af
Consider the simple mirror configuration of Fig. 8.14, As we go from @ to @,
the length of fd? increases while |B| decreases; (8.85) is easily satisfied and we find
instability. The same result is easily obtained for both the 8-pinch and the z-pinch.
Equation (8.85) leads to a very important principle, which we shall not prove
rigorously here. This principle states: Whenever the field lines curve toward the
plasma, the plasma is unstable. Likewise, when field lines curve away from the
plasma, the plasma is interchange stable. Thus, the simple mirror and pinches are
unstable, while the cusp is stable, as shown in Fig. 8.16. Since the field lines curve
away from the plasma the cusp is interchange stable: Furthermore, the magnetic
field is a minimum in the center of the plasma, so that minimum-B is associated
with interchange stability. The instability of the simple mirror has led mirror
designers to look for a minimum-B configuration. They accomplish this by putting
coils, known as Joffe bars, along the axis of the mirror. The net result is a field
configuration that everywhere curves away from the plasma, has a minimum of
|B| in the center, and is MHD stable (Fig. 8.17).
|
Vv
e
(8.85)
8.4 MICROSCOPIC PICTURE OF MHD EQUILIBRIUM
Magnetohydrodynamics is based on a picture of the plasma as a single fluid. Yet
we know that the plasma consists of two species of charged particles. Thus, any
topic in MHD can also be understood by considering the detailed orbits of the
charged particles. In this section, we look at the topic of MHD equilibrium froma
microscopic point of view.Microscopic Picture of MHD Equilibrium 207
Consider a plasma whose density varies smoothly in the £-direction, with mag-
netic field By = Bo(x)2 (Fig. 8.18). Then the MHD equations (8.47) and (8.48)
predict, in the steady state and in component form,
ap.
Pat s,B, (8.86)
and
4x
= a,By = =F J, (8.87)
Suppose that the ions are cold, T; = 0, and the electron temperature T, is a
constant in space. Then at x = 0, (8.86) yields
€ Ny
Wa RT SO. (8.88)
so that our macroscopic picture is one where the plasma pressure is being balanced
by the J * B force. Suppose the ions are cold and very massive; then it is reason-
able to suppose that the current is being contributed by the electrons, even though
the one-fluid equations tell us nothing of the behavior of the individual species. If
this is so, then there must be a mean electron flow speed in the y-direction such
that
= etty (v, ocevon (8.89)
or from (8.88)
cT. 1 an
Yue = (Uydevom = — GRE Ge ‘on (8.90)
which with L,! = (—1/no)(dno/dx) > 0 is
¥ (8.91)
Vac
where v,, is the electron diamagnetic drift speed, and v, is, as usual, the electron
thermal! speed.
Fg, VB
FS
Fig. 8.18 Coordinate system for MHD equilibrium.208 Magnetohydrodynamics
GQ) NY
VM O ,
Fig. 8.19 Single particle picture of MHD equilibria.
EXERCISE Where does the ‘‘diamagnetic” come from?
Now let us look at the single particle picture. The electrons are gyrating about
the magnetic field in the x-y plane. There is no electric field, hence no E X By
drift. There is no curvature of the field lines, hence no curvature drift. There will
be a VB drift but, if we assume a very strong magnetic field, then only a tiny VB is
needed to make V(P + B’/8r) = 0, and we can ignore the VB drift. So where
does the single particle drift come from to make vg, in (8.91)? The fact is that in
terms of guiding centers, there is no drift, but in terms of fluid elements, there is a
drift. Consider the x-y plane, as indicated in Fig. 8.19. Because there are more
particles for x < 0 than for x > 0, a small area of the x-z plane will see more
particles going through to the right than to the left; thus, there is a net flow of
electrons to the right in every fluid element, and a net current to the left. This is
true even though the guiding centers never move!
This concludes our introduction to magnetohydrodynamics. MHD is an ex-
tremely important approximation in plasma physics that is widely used in fusion
plasma physics, solar physics, plasma astrophysics, and energy technology. For
further discussion of MHD sce Refs. [2]-{30].
REFERENCES
[1] G. Schmidt, Physics of High Temperature Plasmas, Academic, New York,
1966.
(2). T. G. Cowling, Magnetohydrodynamics, Interscience, New York, 1957.
[3] R.K.M. Landshoff, ed., Magnetohydrodynamics, Stanford University Press,
Stanford, Calif., 1957.
[4] D. J. Rose and M. Clark, Jr., Plasma and Controlled Fusion, The M.1.T.
Press, Cambridge, Mass., 1961.
(5]_ D. Bershader, ed., Plasma Hydromagnetics, Stanford University Press, Stan-
ford, Calif., 1962.
[(6] W.B. Thompson, An Introduction to Plasma Physics, Pergamon, New York,
1962,
{7} C.L. Longmire, Elementary Plasma Physics, Interscience, New York, 1963.(8)
(9)
(10)
01)
012)
013}
(14)
01s)
(16
07]
18]
19]
(20)
(2
[22]
(23)
[24]
25]
(26)
27
[28]
References 209
S. Gartenhaus, Elements of Plasma Physics, Holt, Rinehart & Winston,
New York, 1964.
P. C. Kendall and C. Plumpton, Magnetohydrodynamics with Hydrodynam-
ics, Vol. 1, Pergamon, New York, 1964.
J. L. Deleroix, Plasma Physics, Wiley, New York, 1965.
E. H. Holt and R. E. Haskell, Foundations of Plasma Dynamics, Macmillan,
New York, 1965
J. A. Shercliff, A Textbook of Magnetohydrodynamics, Pergamon, Oxford,
1965
A, Jeffrey and T. Taniuti, Magnetohydrodynamic Stability and Thermonu-
clear Containment, Academic, New York, 1966.
W. B. Kunkel, ed., Plasma Physics in Theory and Application, McGraw-Hill,
New York, 1966.
M. A. Leontovich, ed., Reviews of Plasma Physics, Vol. 2, Consultants
Bureau, New York, 1966.
A. Simon and W. B. Thompson, eds., Advances in Plasma Physics, Vol. 1,
Interscience, New York, 1968.
T.J.M. Boyd and J. J. Sanderson, Plasma Dynamics, Barnes & Nobel, New
York, 1969.
P.C. Clemmow and J. P. Dougherty, Electrodynamics of Particles and Plas-
mas, Addison-Wesley, Reading, Mass., 1969
H. Cabannes, Theoretical Magnetofluiddynamics, translated by M. Holt and
A. A. Sfeir, Academic, New York, 1970.
N. A. Krall and A. W. Trivelpiece, Principles of Plasma Physics, McGraw-
Hill, New York, 1973.
Course on Instabilities and Confinement in Toroidal Plasmas, Commission of
the European Communities, Luxembourg, 1974.
A.B. Mikhailovskii, Theory of Plasma Instabilities, Vol. 2: Instabilities of an
Inhomogeneous Plasma, translated by J. B. Barbour, Consultants Bureau,
New York, 1974.
M. A. Leontovich, ed., Reviews of Plasma Physics, Vol. 6, Consultants
Bureau, New York, 1975,
D. E. Davis, ed., Pulsed High Beta Plasmas, Pergamon, Oxford, 1976.
G. Bateman, MHD Instabilities, The MIT Press, Cambridge, Mass., 1978.
Plasma Transport, Heating and MHD Theory, Proceedings of the Workshop,
Varenna, Italy, 12-16 September 1977, Pergamon, Oxford, 1978.
Space Plasma Physics: The Study of Solar-System Plasmas, Vol. 1: Reports of
the Study Committee and Advocacy Panels, National Academy of Sciences,
Washington, D.C., 1978.
Space Plasma Physics: The Study of Solar-System Plasmas, Vol. 2: Working
Papers, National Academy of Sciences, Washington, D.C., 1979.210 Mognetohydrodynamics
[29] F. Krause and K.-H. Radler, Mean-Field Magnetohydrodynamics and Dyna-
mo Theory, Pergamon, New York, 1980,
[30] M. A. Leontovich, ed., Reviews of Plasma Physics, Vol. 8, Consultants
Bureau, New York, 1980.
PROBLEMS
8.1 Alfvén Waves
Linearize the ideal MHD equations (8.28) to (8.31) with par = po, B = Bot +
B,V = V$,I = J,8,andk = k2 to obtain the dispersion relation for Alfvén
waves. Compare your result to the result (7.214) from the two-fluid theory, and
explain any differences.
8.2 Magnetosonic Waves
For a cold plasma, linearize the ideal MHD equations (8.28) to (8.31) with py =
Pw, B = By + Bi, V = V9, = J,8, and k = Kf to obtain the dispersion
relation for fast magnetosonic waves. Compare your result to the result (7.227)
from two-fluid theory, and explain any differences.CHAPTER
Discrete Particle Effects
9.1 INTRODUCTION
There are many effects in a plasma that are associated with the discrete nature of
the plasma particles. One of these effects is that of collisions, which is studied in
Chapters t and 3 to 5. A collision is the interaction of one discrete particle with
another discrete particle. There are also discrete particle effects that are due to the
interaction of one discrete particle with the plasma as a whole. For example, a fast
electron moving through a plasma emits Langmuir waves. This phenomena de-
pends on the fact that the fast electron is indeed a discrete particle, but it does not
require that the rest of the plasma be made up of discrete particles; thus, the rest of
the plasma can be treated through the Vlasov approximation. This leads to an
extremely useful approach known as the fest-particle method.
Since the effects to be studied in this chapter are discrete particle effects, it might
have made more sense to study them togethers with the discrete particle collisional
effects. However, it turns out that the test-particle method relies heavily on the
Vlasov dielectric function, Chapter 6 on Vlasov theory taught us many of the
properties of this dielectric function, so that we can now comprehend the results of
the test-particle method more easily.
9.2 DEBYE SHIELDING
Asa first application of the test-particle method, let us calculate the Debye shield-
ing of a test charge g; that moves through a uniform plasma with a constant speed
Yo, starting from position xp = 0 att = 0. For simplicity, we freeze the ions, and
treat the plasma electrons via the Vlasov equation. The only discreteness in the
problem is the test charge. Then Poisson’s equation is212 Discrete Particle Effects
Volx.1) = — 4rp
= = 4relny — f avs. (ev,0] — Arar 80x — vol)
= ne f dv fuls.vsd) = 4g: 5x — vot) OD
where f, = fy + J, and the fy term cancels the ion term. Assuming that the test
charge makes only a small perturbation in the electron density, we can linearize
the Vlasov equation to obtain
Afilant) + VVFi = — F- Ver Vflv) (9.2)
With these two equations, we use Laplace transform techniques to study the initial
value problem. The test charge suddenly appears at x = 0 at 4 = 0, and the
distribution function is initially unperturbed,
Air = 0) = 0 5 (9.3)
We Fourier transform (9.1) and (9.2) in space, and Laplace transform in time, The
Fourier and Laplace transform conventions are stated in Chapter 5. The spatial
Fourier transform of Poisson's equation (9.1) is
— elk) = 4ref aviv.) — 3 even (9.4)
which has the Laplace transform
any!
= kglk,w) = Aref dvfi(kva) + CaS (9.5)
Because the initial value of f, is zero, the Fourier-Laplace transform of the linear-
ized Vlasov equation (9.2) is
(fw + kW Ako) = — Fa tk: WLolWolkw) (9.6)
Solving (9.6) for f, and inserting in (9.5) yield
ane fk Vfl)
“ee fav
iQmy ar
(9.7)
Re(k.w) = — Ako) +
w — kv
w-—k-v
The integrations of the two velocity directions perpendicular to k can be per-
formed, yielding the form
2 (u) ) _ (2° gr
Hetke) [1 - 4 $5 fu 27] -4 Ke Vo (9.8)
where in square brackets we recognize our old friend the Vlasov dielectric function
(6.34). Thus,
iY gr
Rdka)(e a
Next, the Laplace transform is inverted to obtain
(kya) =Debye Shielding 213
Fig. 9.1 Inverse Laplace contour used in calculating the Debye shielding of a test charge.
otk.) = f #2 emigth.o)
a a iQn)" qr
=|, 5e " Facake SE) cae
The pole structure of the integrand is as shown in Fig. 9.1, where for Maxwellian
electrons e(k,w) has among others, the two zeros corresponding to Landau damped
Langmuir waves. For this calculation, we can move the contour downward for
4 > Oas in the calculation of Langmuir waves in Section 6.4. We ignore all of the
transient contributions discussed in Section 6.4 (see Fig. 6.6). Furthermore, let us
ignore the contribution from the Langmuir poles, which will damp away at large
times. Then we pick up only the pole at » = k ¥o, obtaining
OID
Since ¢ is evaluated at the real frequency w = k* Vo, we can use the exact formula
(6.45),
e d, g(u) Pa
ko) = 1 - 35 Pf au oe 77 FE Ae Wlymoe (9-12)
Let us work out several examples.
EXAMPLE A VACUUM
Letting the plasma disappear, w, — 0, we have « = 1. Then
= FT Qwik yt
otk.) = 3a € (9.13)
Performing the inverse Fourier transform we have
est) = f dhe *9tk,0)
= fake = Wet we (9.14)214 Discrete Particle Effects
This integral can be performed in spherical coordinates, letting the k,-axis be in
the direction of x — vot. Then
k +(x ~ vor) = klx — vorleos 6 + °(9.15)
and
ox.) = ae [akon f 9 sin 8 eikix ~tetleos 2 (9.16)
With w = cos 8, du = — sin 6 d@, we have
eae ae ae ‘ dex ~ vt
eeu) = f ak f dy etx ~vtln
ged ( sin(k|x — vot)
=a wns, k (9.17)
[a
| sing
Sa
n/2
or
iieauue 7s
= (9.18)
Thus, we regain the potential due to a point charge in vacuum, moving with
velocity Vo.
EXAMPLE B TEST CHARGE AT REST IN PLASMA (vy, << v,)
For a test charge at rest, or moving very slowly (|Vo] << v,), we expect to regain
the Debye shielding of Chapter | (for motionless ions). Setting w = K+ vy ~ 0,
we have, taking the electrons Maxwellian [see Eq. (6.24)],
fats
aa
~ fai se)
aoe
v2 (9.19)
or
ecko = 0) = 1 + eo (9.20)
which is the “static dielectric function” with fixed ions. The potential is then
Qmylqr _ ny! gr
oh) = Riko =O ~ BER (921)Debye Shielding 215
where we have defined the Debye wave number
k, = ag! (9.22)
without any factor of 27, Then
etx.) = f dk eb etk.o
~ & ftop af eke cos 6
= Sf nk af dosind Gas
- fa rEg we
k sin kx
ea zal ae (9.23)
where the substitution v= cos @ has been used. Because the integrand is even in
k, we can extend the integration to —%° and use contour techniques. We find
(x = |x| > 0)
Lye ksinkx _ eM — =o")
2 I dk ak? rs. ak OF TE ik) (¢
2mi_ | _ik, ~ik,
eee ¢ J ae eit-ikyyx
20 | ik, + om, © ]
ee as (9.24)
so that
x,t) | yy << ye (9.25)
x
which is exactly what we expect for a Debye shielded test particle. Note that this
formula is valid not only for motionless particles, but also for moving particles as
long as vu << v,. (See Refs. [1}{9].)
EXAMPLE C VERY FAST TEST CHARGE (¥, >> v,)
For a very fast test charge, the dielectric function is
(Ka = kev) = 1 $5 pf” du = ee,
fo dude
- 0
at (9.26)
where we have ignored u compared to k - ¥o/k in the denominator. But this is the
same result as for a test charge in vacuum (Example A), so we find6
4
2 ale
2
®
i
? °
3
é
2
4
6
“6 = =2 0 2 4 6
Fig. 9.2. Steady-state equipotential contours, as measured in the frame of a moving test
charge at the origin, with charge g, and speed v9/2 v, = 0.1. Contour labels indicate the
is at zero potential. From Ref. [9]
Distance 2'= (2~08)/ Ry
value of yh,/qy, Unclosed, unlabeled contoui
6 5
vy//Bo, = 0.3
4
2
» j— 0.1
$ 02
:
a)
i 03
7 08
4
Bg eee
“6 4 2 0 2 4
Distance 2°= (2~U98)/ Ry
Fig. 9.3 Same as Fig. 9.2, for vp/2" v, = 0.3.
216+4 4 2 0 2 4 6
Distance 2'= (s—09$)/ A,
Fig. 9.4 Same as Fig. 9.2, for uy/2" v, = 1.0.
Distance ¥=y/ 2,
+ = 2 0 2 4 6
Distance 2° (4-191 hy
Fig. 9.5 Same as Fig. 9.2, for vy/2" v, = 2.0.
217V2, =3
Distance y"=y/2,.
0.1
0.2
0.3
0s
0.8
Fig. 9.6
4 2 a 2
Distance 5= (2—t6)/A,
Same as Fig. 9.2, for v)/2" v, = 3.0.
5
T
wyV2x,* 10
Ye
Distance ¥/
2
Fig. 9.7
218
—4 2 0 2
Distance 2°= (2—%98)/ he
Same as Fig. 9.2, for v/2" v, = 10.0.Fluctuations in Equilibrium 219
a Qr
aD = Sp we PP (9.27)
The plasma does not have time to ‘‘see”’ a fast test charge, and thus does not have
time to respond and shield. (See Refs. [1]-[9].)
We have seen that a fast particle has no shielding, while a slow particle is
completely shielded. A particle moving at intermediate speeds will be partially
shielded. The words fast, slow, and intermediate will depend on which plasma
species we are talking about. Figures 9.2 to 9.7 show the transition from the Debye
shielding of an almost motionless particle [v)/(2“ v,) = 0.1 in Fig. 9.2] to the
almost unshielded behavior of a fast particle [vj/(2% v,) = 10.0 in Fig. 9.7].
In the next section, we continue to exploit the test particle approach, calculating
the equilibrium level of fluctuations in a plasma.
9.3 FLUCTUATIONS IN EQUILIBRIUM
In the previous section, we computed the electrostatic potential in a plasma due to
a single test charge. We found that the rest of the plasma could be treated by the
Vlasov equation.
In this section, we want to use the same ideas to compute the average level of
electric field fluctuations in an equilibrium plasma. We do this by considering each
and every plasma particle as a test charge. Each test charge sees the rest of the
plasma as a Vlasov plasma, and it emits waves satisfying the Cerenkov condition
w = k+ vo, where vy is the velocity of the test charge. Likewise, waves are damped
via Landau damping, which again involves the resonance condition w = k+ Vy =
ku where v, is now the velocity of the particle that is doing the Landau damping,
andu = k-v)/k. Thus, we have a steady-state situation, with waves being emitted
and absorbed, At each point in the plasma, the electric field is fluctuating wildly in
space and time. However, the ensemble averaged electric field energy density is a
constant in space and in time. It is this ensemble averaged electric field energy
density that we wish to calculate,
The process considered here is an example of the principle of detailed balance,
which states that to every emission process, there is a corresponding damping
process, and vice versa, Cerenkov emission of Langmuir waves corresponds to
Landau damping of Langmuir waves. In steady state, these two processes are
balanced; the average rate of wave emission equals the average rate of wave
damping.
Note that here we are using the term “wave” in its most general sense; we have
fluctuations of all frequencies and al! wave numbers, including the normal modes
that we call “Langmuir waves.”
The mathematics involved in this calculation is straightforward. Recali from
Eq, (9.11) that the potential due to a single test charge is written in wave number
space as
(Q2r2y gp ew MD
(kn) = Redkiw =k va) (9.28)220 Discrate Particle Effects
where in the exponent we have specified the orbit vor by the expression Xo(4). Since
F(x,0 = — Pe.) (9.29)
we have
sik * xy)
E(k.) = — iky(k) = — (2n°)"4igy Pain Sy (9.30)
so that
Bot) = f dk eB,
Me xqC2)
k= vo)
ke
Rak,
— (2n*y" grif dk (9.31)
Equation (9.31) gives the electric field at point x due to a particle with orbit x,(#).
Tf we add up the fields at x from all particles in the plasma, and take the ensemble
average, we have
Beat) = f dof csp Kl%.1) fo (%o%0) 9.22)
where fy is the zero order distribution function. The function f, is the probability
density for plasma particles to have velocity vp and position x; thus, the ensemble
average of any plasma property that is caused by the discrete plasma particles is
given by an equation like (9.32).
In a uniform isotropic plasma, there is no preferred direction; therefore we
expect (9.32) to vanish. This indeed bappens.
EXERCISE By performing the x, integration, convince yourself that (9.32)
vanishes.
Next, consider the ensemble average electric field energy density in the plasma,
ae ‘i
W = Za (El) E() (9.33)
Since this is a positive definite quantity, we expect a nonzero result. Taking the
ensemble average in the same way as in (9.32), we have
w= gE fanfan (25) son
“(fa BI [fae SRE] om
where we have used E*(x) = E(x) for the real electric field in the second set of
square brackets. The xy integration yields (27)'6(k — k’) which facilitates the k’
integration; we findFluctuations in Equilibrium 221
2
= f Sf aUecetsetsaeL aesaceaee
W = ge) Mofolvo) | ae = k= volt (9.35)
We can perform the two velocity integrations in the directions perpendicular to k,
and extract a factor of ng, to obtain in the usual fashion
wee maf au g(uyf ak Bake Shor (9.36)
Defining » = ku, we find
me _&(@/k) _
Ais < ( f,; >... >
4, > ty, a Markov process has a probability density such that
PUK n|Xpr Xz +» Xo) = (XulXn-1) (B.23)
where the notation p(a|b) means “the probability density of @ given that b was
true.” Thus, for a Markov process, the probability that x, = 5 depends only on
what the value of x,_; was; it does not depend on what the values of x,..2, X,-3, etc.
were.
There are both discrete and continuous Markov processes. An example of a
discrete Markov process is given by flipping a coin. A trivial example comes if we
give each toss a value x(1,) = x, = + | for a toss of “heads” and a value
xX, = — | for a toss of “tails.” Then x is clearly a Markov process, since p(x,) =
Box, — 1) + 46(x, + 1) does not depend on x,.,, much less on Xp-2, Xp-31 ClC.
A better example of a discrete Markov process is given by defining the random
variableMarkov Processes and Fokker-Planck Equations 273
(1)
Fig. B.6 Any function in nature can be drawn as a smooth curve as shown
X(t) = X, = x (B.24)
where the x; are given by the coin tosses of the previous paragraph. Now X is
clearly a Markov process, whose probability density at f, very definitely depends
on the value of X,_,, but on no previous value.
EXERCISE Calculate p(X,|X,-1) for this example.
To give an example of a continuous Markov process is more difficult, because a
continuous Markov process cannot exist in nature. To see this, consider any
random function that we can draw as a smooth curve, as in Fig. B.6. Now, on the
time scale shown, it appears that x,,, not only depends on x,, but also on x,_). That
is, X,4; not only depends on x,, but also on the derivative of the function
dx(1)/dt},,, Which can be written
dx(t) X, — Xp
dt |=, Ar
Thus, this function is not a Markov process. In fact, no function that is a continu-
ous curve and, therefore no physical function, can be a Markov process.
This does not mean that Markov proccsses cannot be a good approximation to
a physical process. Consider the velocity function of the Brownian particle in the
previous section (Fig. B.7). We have seen that the velocity consists of a rapid
fluctuation due to each molecular collision, together with a slowing down or net
friction force. Thus, on the time scale of molecular collisions, the process is not
Markovian. However, on the much longer time scale of many collision times, the
(B.25)
(ey
'
Fig. B.7 One realization of the velocity of a Brownian particle in a particular direction.274 Appendix B
situation is very nearly Markovian. The Brownian particle is performing a random
walk in velocity space, and soon forgets the details of its orbit near r ; it does,
however, remember its velocity vp at t = 0.
Thus, we consider the process to have three time scales (Fig. B.8): the collision
time 7,, Which is the autocorrelation time of the force SF(t) in the Langevin
equation; the time ¢ after which we may assume to good approximation that fhe
process is Markovian; and the dissipation time v'', We must have At >> 7,; we
shall further assume in this section that At << vy’.
Let us develop some of the mathematical properties of Markov processes. This
development will lead us to the Fokker-Planck equation.
Consider the probability of a sequence of values of the random function x(1).
This is
P(X Xpty + + ++ Xo.X1,Xo) = {probability that, at time fo, the process x(t)
has the value x, and at time t,, x(t) has the
value x,, and... and at time ¢,, x(t) has
the value x,} where, > t)-) > ty»... >
1, > to (B.26)
By the definition (B.23) we can write
P(X rXnery = = >» Xo) = P(X p|XnteXnae - - + + Xo)
XK Xe tXyers oe 0) = PG Opt Xeas es + Xo) (B.27)
The same procedure can now be applied to the last factor on the right of (B.27), so
that
PXnr%nds 6 Xo) S P(XnilXn-2)O( nas» «+ + Xo) (B.28)
and so on until we have finally, for a Markov process,
et)
%
t
Hi Collision time, not Markovian
[at Many collision times, almost Markovian
|-___, _1 Dissipation time
»
B.8 Three time scales of Brownian motion.Markov Processes and Fokker-Planck Equations 275
PX Xmty%ay oe + 9 Xo) = (XI Xn |Xn-2)
++ o(%21%1)e(% 1X0)a(%o) (B.29)
By elementary considerations it must also be true that
PX nr XnevXnezs + +s XyXo) = MKa%eay e+» « XilXo)O(%o) (B.30)
Comparing (B.30) and (B.29) we find
PU XpXyiy 0 Xe1%o)
= A XnlXn-1)O(%eal%r-2) + + - OC 1%0) (B31)
In particular, we can choose n = 2 to obtain
(2X1 1X0) = (21% )o(%1 1X0) (B.32)
Let us row integrate this expression over all possible x, to obtain
p(%2|Xo) =f de (%2,%1 1X9) (B.33)
elaalre) = fd, ebrate Grn) (B34)
which is the Chapman-Kolmogorov equation, or Smoluchowsky equation (8].
Suppose we identify x, with time ¢ and x, as x(t + Ar). Suppose we further
assume that
or
AX) = plxst = to) = (x — Xo) (B.35)
Then we can drop the references to xp in (B.34), and write
P(xX2/Xo) = p(x.t + Ar) (B.36)
that is, x, is now denoted by x, and
(x, 1X0) = p(X) (B.37)
We can also change the notation of p(x,|x;); with the definition
Ax=x— x, (B.38)
we can write
Cx|%) = pl(x,t + Atlx — Ax,t)
u(Ax,t + At|x — Ax,t) (B.39)
where the transition probability is defined by (B.39); gives the probability that
at time ¢ + At, the random process has made a jump of Ax from its previous
value x — Ax at time J.
With these notational! changes, we can rewrite (B.34) as
p(x,t + At) = f danwars + At|x — Ax,Op(x — Ax,t) (B.40)276 Appendix B
The value x appears on the right of (B.40) only in the combination x — Ax. Thus,
if we assume that all of the important physics happens for small Sx, then we can
make a Taylor series expansion on the right of (B.40), obtaining
p(x,t + Ax) = faax s (rant
oe
|
[wane + At|x — Ax,t)p(x — ax] HIE}
or
i = (=Ax) a!
p(xt + An) = faax) = 7
X (W(dx,t + Arlx,n)p(x,1)) (B.41)
If the infinite sum converges, and if we can interchange the summation and inte-
gration, then we can write
sev a
p(x,t + dt) = = ested
X (olen f danianiware + adlx0] (B.42)
The quantity given by the Ax integration is just the expectation value or ensemble
average of (Ax)',
(4x) = f qanadiwars + Atlx,n) (B.43)
which is itself a function of x,t through y. Equation (B.42) becomes
2 (1 a! ,
post + At) = > ar [olsX(O0)' 4] (B.44)
= 7
Moving the / = 0 term to the left side, and dividing by As, we have
p(x, + At) — plxt) _ ¥ Cy a
at fe dt ax!
[ele X(Axy1)] (B45)
We next take the limit as At — “0”. This means that we Jet Ar become very small,
much smaller than any macroscopic time scale (e.g., »"!). However, At cannot
really go to zero, because this development has assumed that Ar is large enough to
justify the Markovian assumption. Thus, the left side of (B.45) becomes
plxt + A) = ox.) _ dolx.t)
at ar (B.46)
lim
amo"
where the time derivative refers to macroscopic time. Equation (B.45} becomesMarkov Processes and Fokker-Planck Equations 277
ae (dx)
= 2 (-l! oy [um ne oss] (B.47)
a = !
Defining the diffusion coefficients
i (dx)!
Dx.) = Min te a) (B.48)
Equation (B.47) is
Seed) Be 1 © De%s.npts, ) (B.49)
If we keep only the first two a on the right of (B.49), we have
So 2 ipo & pe
BP Gy DOOMED] + Fs DMN] | (B50)
which is the well-known Fokker-Planck equation [9]
For Brownian motion, the random variable x is replaced by the particle velocity
u(f). We shall leave it as an exercise to determine the diffusion coefficients D'!(v,t)
and D?y,1).
EXERCISE Use the results of the previous section to evaluate the coefficients in
Eq. (B.50) for Brownian motion. Show that D'"(v,t) = — vu, and D?(u,1) =
vT/M, so that the Fokker-Planck equation associated the Langevin equa-
tion of Brownian motion is
8p(v,t) oT #
at ay (P) + a aye (B.51)
EXERCISE Use the results of the previous section to show that D"(v,t) ~ At
and, thus, vanishes as Ar — “0”
We can now understand why we are able to write the Lenard-Balescu equation
in the form of a Fokker-Planck equation,
ast) —
1 fe
at WAN + 7 Vs, Vy BN (B.52)
Because the derivation of Lenard-Balescu assumed g(1,2) << f,(1)f,(2), we have
effectively limited ourselves to small angle two-body collisions. The quantity /(v,,t)
may be thought of as the probability density of particles in velocity space. Thus,
(¥,,1) is changing slowly on the time scale for a two-body collision. All of these
features are precisely those assumed in the derivation of the Fokker-Planck equa-
tien, It should come as no surprise to us that the Lenard-Balescu equation can be
written in the form of the Fokker—Planck equation. The coefficient A in (B.52) is
called the coefficient of dynamic friction, and plays the same role as vv in the
Fokker-Planck equation (B.51) for Brownian motion. It represents the slowing278 Appendix B
down of a particle due to many small angle Coulomb collisions, Likewise, the
coefficient B in (B.52) is called the diffusion coefficient, and plays the same role as
vT/M in (B.51). It represents the diffusion of the plasma particles in velocity space
due to many small angle collisions.
In the steady state, a typical particle is suffering dynamic friction plus diffusion;
the net effect is to produce a Maxwellian. This is just as true in a plasma as itis for
a Brownian particle.
In addition to the stated references, sources for this appendix include the book
by Stratonovich [10] and the ageless and excellent article by Chandrasekhar (11]
REFERENCES
[1] R. Brown, Phil. Mag., 4, 161 (1928).
[2] R. Brown, Ann. Phys. Chem., 14, 294 (1928).
[3] M. P. Langevin, C. R. Acad. Sci. Paris, 146, 530 (1908).
[4] A. Einstein, Ann. Phys., 17, 549 (1905).
[5] A. Einstein, Ann. Phys., 19, 289 (1906).
[6] A. Einstein, Ann. Phys., 19, 371 (1906).
[7] D. K. C. MacDonald, Noise and Fluctuations: An Introduction, Wiley, New
York, 1962.
[8] _M. von Smotuchowsky, Ann. Phys., 21, 756 (1906).
[9] Lord Rayleigh, Nature, 72, 318 (1905).
[10] R. L. Stratonovich, Topics in the Theory of Random Noise, Gordon and
Breach, New York, 1963, Vol. 1.
(11) S. Chandrasekhar, Rev. Mod. Phys., 15, 1 (1943)APPENDIX
Pedestrian’s Guide to
Complex Variables
Many parts of this book make use of the basic results of the theory of complex
variables [1, 2]. For the benefit of readers who have not yet studied this subject in
detail, or who have studied it long enough ago to have forgotten it, these basic
results are summarized here.
The most useful result is the residue theorem, which states that the integral in a
counterclockwise direction around a closed curve is 27/ times the sum of the
residues. If the integrand is of the form f(z(z — 2)"!, the residue at the simple
pole z = 2 is f(z9). For example, consider the integral
aa dz
where the integration is along the real z-axis, and @ > 0. The integration can be
closed by a large semicircle at infinity, since the contribution from the semicircle is
~ lim (7R/R*) = lim (a/R) = 0
Roe Row
The semicircle can occur in either the upper-half z-plane or the lower-half z-plane
(Fig. C.1). Writing (C.1) as
I -¢ Seas 2 EE
( + ia\z — ia)
we close the contour downward, changing the sign on the result because this is in
the clockwise direction, to obtain (only the pole at z = — ia is enclosed)
(€.2)
z— ia
a2 (C3)
ia
I = (-2n/)280 Appendix C
Fig. C.1 Integration contour in the z-plane.
EXERCISE Obtain the same result by closing upward.
The solution of many ordinary and partial differential equations is facilitated by
Fourier and Laplace transformation. The Fourier uransform conventions used in
this book, stated in Chapter 5, are for functions of one spatial dimension x,
Ho =f" & seep cinn (C4)
with inverse transform
f(x) = J ” dk (ke) exp (ikx) (C5)
The x and k integrations are along the real axes. The Laplace transform conven-
tions for functions of time f are
Say = [7 at f(0) exp (iat) (C6)
F
with inverse transform
dw
Ko =f FE fleyexp cian) cn
where the Laptace contour L is a horizontal line in the complex w-plane that must
pass above all singularities of f(w). The Laplace transform (C.6) can be considered
as the Fourier transform of the function 7({#) such that 7(t) = f(t) fort > 0 and
F(t) = 0 for t < 0. Then for + < 0 the inverse Laplace transform (C.7) can be
closed upward [since exp (—iwt) ~ exp (—|t|w,) ~ 0 for w; + +], yielding
F() = O fort < 0 since the Laplace contour passes above all singularities of f(w).
Consider the solution of the differential equation
q
4 = af (C.8)Pedestrian’s Guide to Complex Variables 281
with f(¢ = 0) = fo. The Laplace transform of the left side is
£ a ane exp (fet) = f(1) exp (ia)|~
- taf a S(t) exp (dwt) = f(1) exp (iat)|” — iw f(o) (C9)
°
Without knowing the function /(/), we can only say that this integral is defined for
w; large enough, for only then is f(t) exp (‘t)| -.. equal to zero. How large is large
enough remains to be seen. For large enough w,, Eq. (C.8) has the Laplace
transform
- ft = 0) ~ iwf(w) = a fla) (C.10)
so that
_ ft = 9
BAC) Beaerremnera (C1)
The inverse transform is
gy =f % = op cian (C.12)
2m wie
where the contour must be placed high enough in the wplane so that f(w) is
defined (the shaded region in Fig. C.2). Once the contour is drawn in the shaded
region of Fig. C.2, it can be moved around only if f(a) is analytically continued to
the remainder of the complex w-plane. An ana/ytic function is one that is differen-
tiable (the derivative in the complex plane at the point z does not depend on which
direction the point is approached from). The analytic continuation of a simple
function like f(w) in (C.11) from the shaded region in Fig. C.2 (the only region
where (C.10) is defined} to the entire orplane is easy; it is just the function
so) = 9 (C.13)
Fig. C.2- Inverse Laplace contour must be drawn initially high in the @-plane,282 Appendix C
itself. This function is now analytic in the entire w-plane except at the point
w = ia, The contour in (C.12) can now be closed by a large semicircle in the
lower-half w-plane, since f(w) is defined everywhere. The result is
2rilé
ray = - =O. Fa = oyexp (an
7
= f(t = 0) exp (at) (C.14)
which is the desired result. In retrospect, we can now see that (C.9) converges for
@, > a. This is why the inverse Laplace contour must be drawn above ail singular-
ities of f(w) in the w-plane.
Analytic continuation is not always quite as simple as in (C.13). Consider
il dx
r= f" ats (C)5)
defined for z,; > Owitha > 0; the integration is along the real x-axis. Closing the
contour either up or down (Fig. C.3), we find
1
a zti (C16)
f(z) =
for z, > 0. We canmot use (C.15) as the analytic continuation of f(z) for z; < 0,
for then (C.15) yields
_ 1 2n
$0) Tia F+e@ ea
for z; < 0. The function f(z) defined by (C.16) and (C.17) is discontinuous at
z; = 0 and so is not analytic. In order to properly analytically continue (C.15), one
must subtract the extra term in (C.17) that Jeads to the discontinuity, and write
ft dx Qi
sea) =f @tae@-a tte (C18)
Fig. C.3_ Evaluation of f(z) for z, > 0 in (C.15).Pedestrian's Guide to Complex Verlebles 263
Fig. C.4 Integration contour for z, < 0 that gives the proper analytic continuation of
(C.15).
for z, < 0. The combination (C.15) for z; > 0 and (C.18) for z, < 0 is now an
analytic function everywhere except at the pole of (C.16),z = — ia. Alternatively,
one can deform the contour in (C.15) as shown in Fig. C.4 for z; < 0, and write
ix
@
F(a) =f @+a@—-_ (C.19)
The contour C’ is as shown in Fig. C.4 for z, < 0, and is along the real x-axis for
z, > 0. This gives the same result as the form (C.18) for z; < 0.
A useful formula for Fourier transformation is
f & exp (-ikx) = &(k) (C.20)
This formula can be demonstrated by multiplying each side by an arbitrary func-
tion f(k) and integrating over all k. The right side yields f(k = 0), while the left
side is
Ji ak p(k) f £ exp (“ikx) “f ZS dk f(k) exp (-ikx)
=[" $1
= f(k = 0) (C.21)
where it has been assumed that the order of integrations can be reversed. Since the
right and left sides of (C.20) yield the same result, the identification (C.20) must be
correct.
Another useful formula concerns integrals of the form
aa a 1
= lim Sie x-asim (C22)284 Appendix C
atin
Fig. C.5__ Integration contour leading to the Plemelj formulas
where the integral is along the real x-axis, anda > 0. For the lower sign, the pole
is at x = a + ilm|, and the integral can be performed by slightly deforming the
contour as shown in Fig. C.5. This leads to
r= Pf” ae 4 + wi (€.23)
where the semicircle in Fig. C.5 contributes half of 2a/, and where
Pf. aque et +S (C.24)
Formally, one writes
: ! es 1 ne
lim soot 7? (gAg) + rte a) (C25)
which when integrated over x yields (C.23). For the upper sign in (C.22), the pole
approaches the integration contour from below, the integration contour must be
deformed upward rather than downward, and the sign of the imaginary contri-
bution changes. The general formula is finally
1
=p{—- 16x —
praeaa 7? (gAq) = me 2) (C.26)
lim.
0
Other properties of complex variables are explored throughout the book.
REFERENCES
(11. S. Sokolnikoff and R. M. Redheffer, Mazhematics of Physics and Modern
Engineering, 2nd edition, McGraw-Hill, New York, 1966.
(2], G. Arfken, Mathematical Methods for Physicists, Academic, New York, 1970.APPENDIX D
Vector and Tensor
Identities
The following vector and tensor identities are useful in the study of plasma physics
a).
A+ (BX C) = (A X B)-C =B-(C X A) = (BX C)-A
= C-+(A XB) = (CX A)-B (D.1)
A x (BX C) = (A: OB — (A BIC (D.2)
(A X B)+(C X D) = (A+ CB -D) — (A+ DYB-C) (D.3)
VUE) = fVg + evs (D4)
\ VGA) S=SVA FAVS (D.5)
VX (fA) =fVKATUSXA (D.6)
V-(A XB) =B+(V x A) — A+(V XB) (D.7)
Vv x (A X B) = A(V-B) — B(V-A) + (B- VA —(A- VB (D.8)
ViA-B) = A X (VX B) + BX (VX A) + (A> VB + (B- VA (D.9)
VA = WV+A)— Vx (VX A) (D.10)
vx vf=0 (D.11)
V-(V KX A=0 (D.12)
V+ (BA) = A(V-B) + (B+ VA (D.13)
VisSH=V/A-T+ys/0-F (D.14)
REFERENCE
[1] D. L. Book, Plasma Formulary, Naval Research Laboratory, Washington,
D.C., 1980.